23.05.2014 Views

Springer Lecture Notes in Physics 716

Springer Lecture Notes in Physics 716

Springer Lecture Notes in Physics 716

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

7<br />

Field-Theory Approaches<br />

to Nonequilibrium Dynamics<br />

U. C. Täuber<br />

Department of <strong>Physics</strong>, Center for Stochastic Processes <strong>in</strong> Science and Eng<strong>in</strong>eer<strong>in</strong>g<br />

Virg<strong>in</strong>ia Polytechnic Institute and State University<br />

Blacksburg, Virg<strong>in</strong>ia 24061-0435, USA<br />

tauber@vt.edu<br />

It is expla<strong>in</strong>ed how field-theoretic methods and the dynamic renormalisation<br />

group (RG) can be applied to study the universal scal<strong>in</strong>g properties of systems<br />

that either undergo a cont<strong>in</strong>uous phase transition or display generic scale <strong>in</strong>variance,<br />

both near and far from thermal equilibrium. Part 1 <strong>in</strong>troduces the<br />

response functional field theory representation of (nonl<strong>in</strong>ear) Langev<strong>in</strong> equations.<br />

The RG is employed to compute the scal<strong>in</strong>g exponents for several universality<br />

classes govern<strong>in</strong>g the critical dynamics near second-order phase transitions<br />

<strong>in</strong> equilibrium. The effects of reversible mode-coupl<strong>in</strong>g terms, quench<strong>in</strong>g<br />

from random <strong>in</strong>itial conditions to the critical po<strong>in</strong>t, and violat<strong>in</strong>g the<br />

detailed balance constra<strong>in</strong>ts are briefly discussed. It is shown how the same<br />

formalism can be applied to nonequilibrium systems such as driven diffusive<br />

lattice gases. Part 2 describes how the master equation for stochastic particle<br />

reaction processes can be mapped onto a field theory action. The RG is<br />

then used to analyse simple diffusion-limited annihilation reactions as well as<br />

generic cont<strong>in</strong>uous transitions from active to <strong>in</strong>active, absorb<strong>in</strong>g states, which<br />

are characterised by the power laws of (critical) directed percolation. Certa<strong>in</strong><br />

other important universality classes are mentioned, and some open issues are<br />

listed.<br />

7.1 Critical Dynamics<br />

Field-theoretic tools and the renormalisation group (RG) method have had<br />

a tremendous impact <strong>in</strong> our understand<strong>in</strong>g of the universal power laws that<br />

emerge near equilibrium critical po<strong>in</strong>ts (see, e.g., Refs. [1–6]), <strong>in</strong>clud<strong>in</strong>g the<br />

associated dynamic critical phenomena [7,8]. Our goal here is to similarly describe<br />

the scal<strong>in</strong>g properties of systems driven far from thermal equilibrium,<br />

which either undergo a cont<strong>in</strong>uous nonequilibrium phase transition or display<br />

generic scale <strong>in</strong>variance. We are then confronted with captur<strong>in</strong>g the (stochastic)<br />

dynamics of the long-wavelength modes of the “slow” degrees of freedom,<br />

U.C. Täuber: Field-Theory Approaches to Nonequilibrium Dynamics, Lect. <strong>Notes</strong> Phys. <strong>716</strong>,<br />

295–348 (2007)<br />

DOI 10.1007/3-540-69684-9 7 c○ <strong>Spr<strong>in</strong>ger</strong>-Verlag Berl<strong>in</strong> Heidelberg 2007


296 U. C. Täuber<br />

namely the order parameter for the transition, any conserved quantities, and<br />

perhaps additional relevant variables. In these lecture notes, I aim to briefly<br />

describe how a representation <strong>in</strong> terms of a field theory action can be obta<strong>in</strong>ed<br />

for (1) general nonl<strong>in</strong>ear Langev<strong>in</strong> stochastic differential equations [8, 9]; and<br />

(2) for master equations govern<strong>in</strong>g classical particle reaction–diffusion systems<br />

[10–12]. I will then demonstrate how the dynamic (perturbative) RG<br />

can be employed to derive the asymptotic scal<strong>in</strong>g laws <strong>in</strong> stochastic dynamical<br />

systems; to <strong>in</strong>fer the upper critical dimension d c (for dimensions d ≤ d c ,<br />

fluctuations strongly affect the universal scal<strong>in</strong>g properties); and to systematically<br />

compute the critical exponents as well as to determ<strong>in</strong>e further universal<br />

properties <strong>in</strong> various <strong>in</strong>trigu<strong>in</strong>g dynamical model systems both near and far<br />

from equilibrium. (For considerably more details, especially on the more technical<br />

aspects, the reader is referred to Ref. [13].)<br />

7.1.1 Cont<strong>in</strong>uous Phase Transitions and Critical Slow<strong>in</strong>g Down<br />

The vic<strong>in</strong>ity of a critical po<strong>in</strong>t is characterised by strong correlations and large<br />

fluctuations. The system under <strong>in</strong>vestigation is then behav<strong>in</strong>g <strong>in</strong> a highly cooperative<br />

manner, and as a consequence, the standard approximative methods of<br />

statistical mechanics, namely perturbation or cluster expansions that assume<br />

either weak <strong>in</strong>teractions or short-range correlations, fail. Upon approach<strong>in</strong>g<br />

an equilibrium cont<strong>in</strong>uous (second-order) phase transition, i.e., for |τ| ≪1,<br />

where τ =(T − T c )/T c measures the deviation from the critical temperature<br />

T c , the thermal fluctuations of the order parameter S(x) (which characterises<br />

the different thermodynamic phases, usually chosen such that the thermal<br />

average 〈S〉 = 0 vanishes <strong>in</strong> the high-temperature “disordered” phase) are, <strong>in</strong><br />

the thermodynamic limit, governed by a diverg<strong>in</strong>g length scale<br />

ξ(τ) ∼|τ| −ν . (7.1)<br />

Here, we have def<strong>in</strong>ed the correlation length via the typically exponential<br />

decay of the static cumulant or connected two-po<strong>in</strong>t correlation function<br />

C(x) =〈S(x) S(0)〉 −〈S〉 2 ∼ e −|x|/ξ ,andν denotes the correlation length<br />

critical exponent. AsT → T c , ξ →∞, which entails the absence of any characteristic<br />

length scale for the order parameter fluctuations at criticality. Hence<br />

we expect the critical correlations to follow a power law C(x) ∼|x| −(d−2+η)<br />

<strong>in</strong> d dimensions, which def<strong>in</strong>es the Fisher exponent η. The follow<strong>in</strong>g scal<strong>in</strong>g<br />

ansatz generalises this power law to T ≠ T c , but still <strong>in</strong> the vic<strong>in</strong>ity of the<br />

critical po<strong>in</strong>t,<br />

C(τ,x) =|x| −(d−2+η) ˜C± (x/ξ) , (7.2)<br />

with two dist<strong>in</strong>ct regular scal<strong>in</strong>g functions ˜C + (y) forT>T c and ˜C − (y) for<br />

T


7 Field-Theory Approaches to Nonequilibrium Dynamics 297<br />

As we will see <strong>in</strong> Subsect. 7.1.5, there are only two <strong>in</strong>dependent static<br />

critical exponents. Consequently, it must be possible to use the static scal<strong>in</strong>g<br />

hypothesis (7.2) or (7.3), along with the def<strong>in</strong>ition (7.1), to express the<br />

exponents describ<strong>in</strong>g the thermodynamic s<strong>in</strong>gularities near a second-order<br />

phase transition <strong>in</strong> terms of ν and η through scal<strong>in</strong>g laws. For example, the<br />

order parameter <strong>in</strong> the low-temperature phase (τ < 0) is expected to grow<br />

as 〈S〉 ∼ (−τ) β . Let us consider Eq. (7.2) <strong>in</strong> the limit |x| → ∞.Inorder<br />

for the |x| dependence to cancel, ˜C± (y) ∝|y| d−2+η for large |y|, and<br />

therefore C(τ,|x| → ∞) ∼ ξ −(d−2+η) ∼ |τ| ν(d−2+η) . On the other hand,<br />

C(τ,|x| →∞) →−〈S〉 2 ∼−(−τ) 2β for T < T c ; thus we identify the order<br />

parameter critical exponent through the hyperscal<strong>in</strong>g relation<br />

β = ν (d − 2+η) . (7.4)<br />

2<br />

Let us next consider the isothermal static susceptibility χ τ , which accord<strong>in</strong>g to<br />

the equilibrium fluctuation–response theorem is given <strong>in</strong> terms of the spatial<br />

<strong>in</strong>tegral of the correlation function C(τ,x): χ τ (τ) =(k B T ) −1 lim q→0 C(τ,q).<br />

But Ĉ±(p) ∼ |p| 2−η as p → 0 to ensure nons<strong>in</strong>gular behaviour, whence<br />

χ τ (τ) ∼ ξ 2−η ∼|τ| −ν(2−η) , and upon def<strong>in</strong><strong>in</strong>g the associated thermodynamic<br />

critical exponent γ via χ τ (τ) ∼|τ| −γ , we obta<strong>in</strong> the scal<strong>in</strong>g relation<br />

γ = ν (2 − η) . (7.5)<br />

The scal<strong>in</strong>g laws (7.2), (7.3) as well as scal<strong>in</strong>g relations such as (7.4) and<br />

(7.5) can be put on solid foundations by means of the RG procedure, based on<br />

an effective long-wavelength Hamiltonian H[S], a functional of S(x), that captures<br />

the essential physics of the problem, namely the relevant symmetries <strong>in</strong><br />

order parameter and real space, and the existence of a cont<strong>in</strong>uous phase transition.<br />

The probability of f<strong>in</strong>d<strong>in</strong>g a configuration S(x) at given temperature<br />

T is then given by the canonical distribution<br />

P eq [S] ∝ exp (−H[S]/k B T ) . (7.6)<br />

For example, the mathematical description of the critical phenomena for an<br />

O(n)-symmetric order parameter field S α (x), with vector <strong>in</strong>dex α =1,...,n,<br />

isbasedontheLandau–G<strong>in</strong>zburg–Wilson functional [1–6]<br />

∫<br />

H[S] = d d x ∑ [ r<br />

2 [Sα (x)] 2 + 1 2 [∇Sα (x)] 2<br />

α<br />

+ u 4! [Sα (x)] ∑ ]<br />

2 [S β (x)] 2 − h α (x) S α (x) , (7.7)<br />

β<br />

where h α (x) is the external field thermodynamically conjugate to S α (x),<br />

u>0 denotes the strength of the nonl<strong>in</strong>earity that drives the phase transformation,<br />

and r is the control parameter for the transition, i.e., r ∝ T − T 0 c ,


298 U. C. Täuber<br />

where Tc<br />

0 is the (mean-field) critical temperature. Spatial variations of the<br />

order parameter are energetically suppressed by the term ∼ [∇S α (x)] 2 ,and<br />

the correspond<strong>in</strong>g positive coefficient has been absorbed <strong>in</strong>to the fields S α .<br />

We shall, however, not pursue the static theory further here, but <strong>in</strong>stead<br />

proceed to a full dynamical description <strong>in</strong> terms of nonl<strong>in</strong>ear Langev<strong>in</strong> equations<br />

[7, 8]. We will then formulate the RG with<strong>in</strong> this dynamic framework,<br />

and there<strong>in</strong> demonstrate the emergence of scal<strong>in</strong>g laws and the computation<br />

of critical exponents <strong>in</strong> a systematic perturbative expansion with respect to<br />

the deviation ɛ = d − d c from the upper critical dimension.<br />

In order to construct the desired effective stochastic dynamics near a critical<br />

po<strong>in</strong>t, we recall that correlated region of size ξ become quite large <strong>in</strong> the<br />

vic<strong>in</strong>ity of the transition. S<strong>in</strong>ce the associated relaxation times for such clusters<br />

should grow with their extent, one would expect the characteristic time<br />

scale for the relaxation of the order parameter fluctuations to <strong>in</strong>crease as well<br />

as T → T c , namely<br />

t c (τ) ∼ ξ(τ) z ∼|τ| −zν , (7.8)<br />

which <strong>in</strong>troduces the dynamic critical exponent z that encodes the critical<br />

slow<strong>in</strong>g down at the phase transition; usually z ≥ 1. S<strong>in</strong>ce the typical relaxation<br />

rates therefore scale as ω c (τ) =1/t c (τ) ∼|τ| zν , we may utilise the<br />

static scal<strong>in</strong>g variable p = q ξ to generalise the crucial observation (7.8) and<br />

formulate a dynamic scal<strong>in</strong>g hypothesis for the wavevector-dependent dispersion<br />

relation of the order parameter fluctuations [14, 15],<br />

ω c (τ,q) =|q| z ˆω ± (q ξ) . (7.9)<br />

We can then proceed to write down dynamical scal<strong>in</strong>g laws by simply<br />

postulat<strong>in</strong>g the additional scal<strong>in</strong>g variables s = t/t c (τ) orω/ω c (τ,q). For<br />

example, as an immediate consequence we f<strong>in</strong>d for the time-dependent mean<br />

order parameter<br />

〈S(τ,t)〉 = |τ| β Ŝ(t/t c ) , (7.10)<br />

with Ŝ(s →∞) = const., but Ŝ(s) ∼ s−β/zν as s → 0 <strong>in</strong> order for the τ<br />

dependence to disappear. At the critical po<strong>in</strong>t (τ = 0), this yields the powerlaw<br />

decay 〈S(t)〉 ∼t −α , with<br />

α = β<br />

zν = 1 (d − 2+η) . (7.11)<br />

2 z<br />

Similarly, the scal<strong>in</strong>g law for the dynamic order parameter susceptibility (response<br />

function) becomes<br />

χ(τ,q,ω)=|q| −2+η ˆχ ± (q ξ,ω ξ z ) , (7.12)<br />

which constitutes the dynamical generalisation of Eq. (7.3), for χ(τ,q, 0) =<br />

(k B T ) −1 C(τ,q). Upon apply<strong>in</strong>g the fluctuation–dissipation theorem, valid <strong>in</strong><br />

thermal equilibrium, we therefrom obta<strong>in</strong> the dynamic correlation function


C(τ,q,ω)= 2k BT<br />

ω<br />

7 Field-Theory Approaches to Nonequilibrium Dynamics 299<br />

Im χ(τ,q,ω)=|q| −z−2+η Ĉ ± (q ξ,ω ξ z ) , (7.13)<br />

and for its Fourier transform <strong>in</strong> real space and time,<br />

∫ d d ∫<br />

q dω<br />

C(τ,x,t)=<br />

ei(q·x−ωt) (2π) d C(τ,q,ω)=|x| −(d−2+η) ˜C± (x/ξ, t/ξ z ) ,<br />

2π<br />

(7.14)<br />

which reduces to the static limit (7.2) if we set t =0.<br />

The critical slow<strong>in</strong>g down of the order parameter fluctuations near the<br />

critical po<strong>in</strong>t provides us with a natural separation of time scales. Assum<strong>in</strong>g<br />

(for now) that there are no other conserved variables <strong>in</strong> the system, which<br />

would constitute additional slow modes, we may thus resort to a coarse-gra<strong>in</strong>ed<br />

long-wavelength and long-time description, focus<strong>in</strong>g merely on the order parameter<br />

k<strong>in</strong>etics, while subsum<strong>in</strong>g all other “fast” degrees of freedom <strong>in</strong> random<br />

“noise” terms. This leads us to a mesoscopic Langev<strong>in</strong> equation for the slow<br />

variables S α (x,t)oftheform<br />

∂S α (x,t)<br />

= F α [S](x,t)+ζ α (x,t) . (7.15)<br />

∂t<br />

In the simplest case, the systematic force terms here just represent purely<br />

relaxational dynamics towards the equilibrium configuration [16],<br />

F α [S](x,t)=−D<br />

δH[S]<br />

δS α (x, t) , (7.16)<br />

where D represents the relaxation coefficient, and H[S] is aga<strong>in</strong> the effective<br />

Hamiltonian that governs the phase transition, e.g. given by Eq. (7.7). For<br />

the stochastic forces we may assume the most convenient form, and take them<br />

to simply represent Gaussian white noise with zero mean, 〈ζ α (x,t)〉 = 0, but<br />

with their second moment <strong>in</strong> thermal equilibrium fixed by E<strong>in</strong>ste<strong>in</strong>’s relation<br />

〈<br />

ζ α (x,t) ζ β (x ′ ,t ′ ) 〉 =2k B TDδ(x − x ′ ) δ(t − t ′ ) δ αβ . (7.17)<br />

As can be verified by means of the associated Fokker–Planck equation for the<br />

time-dependent probability distribution P[S, t], Eq. (7.17) guarantees that<br />

eventually P[S, t →∞] →P eq [S], the canonical distribution (7.6). The stochastic<br />

differential equation (7.15), with (7.16), the Hamiltonian (7.7), and<br />

the noise correlator (7.17), def<strong>in</strong>e the relaxational model A (accord<strong>in</strong>g to the<br />

classification <strong>in</strong> Ref. [7]) for a nonconserved O(n)-symmetric order parameter.<br />

If, however, the order parameter is conserved, wehavetoconsidertheassociated<br />

cont<strong>in</strong>uity equation ∂ t S α +∇·J α = 0, where typically the conserved<br />

current is given by a gradient of the field S α : J α = −D ∇S α + ...; as a consequence,<br />

the order parameter fluctuations will relax diffusively with diffusion<br />

coefficient D. The ensu<strong>in</strong>g model B [7,16] for the relaxational critical dynamics<br />

of a conserved order parameter can be obta<strong>in</strong>ed by replac<strong>in</strong>g D →−D ∇ 2 <strong>in</strong><br />

Eqs. (7.16) and (7.17). In fact, we will henceforth treat both models A and B


300 U. C. Täuber<br />

simultaneously by sett<strong>in</strong>g D → D (i∇) a , where a =0anda = 2 respectively<br />

represent the nonconserved and conserved cases. Explicitly, we thus obta<strong>in</strong><br />

∂S α (x,t)<br />

∂t<br />

= −D (i∇) a δH[S]<br />

δS α (x,t) + ζα (x,t)<br />

= −D (i∇) a[ r − ∇ 2 + u ∑<br />

[S β (x)] 2] S α (x,t)<br />

6<br />

β<br />

+D (i∇) a h α (x,t)+ζ α (x,t) , (7.18)<br />

with<br />

〈<br />

ζ α (x,t) ζ β (x ′ ,t ′ ) 〉 =2k B TD(i∇) a δ(x − x ′ ) δ(t − t ′ ) δ αβ . (7.19)<br />

Notice already that the presence or absence of a conservation law for the order<br />

parameter implies different dynamics for systems described by identical static<br />

behaviour. Before proceed<strong>in</strong>g with the analysis of the relaxational models,<br />

we remark that <strong>in</strong> general there may exist additional reversible contributions<br />

to the systematic forces F α [S], see Subsect. 7.1.6, and/or dynamical modecoupl<strong>in</strong>gs<br />

to additional conserved, slow fields, which effect further splitt<strong>in</strong>g<br />

<strong>in</strong>to several dist<strong>in</strong>ct dynamic universality classes [6, 7, 13].<br />

Let us now evaluate the dynamic response and correlation functions <strong>in</strong> the<br />

Gaussian (mean-field) approximation <strong>in</strong> the high-temperature phase. To this<br />

end, we set u = 0 and thus discard the nonl<strong>in</strong>ear terms <strong>in</strong> the Hamiltonian<br />

(7.7) as well as <strong>in</strong> Eq. (7.18). The ensu<strong>in</strong>g Langev<strong>in</strong> equation becomes l<strong>in</strong>ear <strong>in</strong><br />

the fields S α , and is therefore readily solved by means of Fourier transforms.<br />

Straightforward algebra and regroup<strong>in</strong>g some terms yields<br />

[<br />

−iω + Dq<br />

a ( r + q 2)] S α (q,ω)=Dq a h α (q,ω)+ζ α (q,ω) . (7.20)<br />

With 〈ζ α (q,ω)〉 = 0, this gives immediately<br />

χ αβ<br />

(q,ω)〉<br />

0 (q,ω)=∂〈Sα ∂h β (q,ω) ∣ = Dq a G 0 (q,ω) δ αβ , (7.21)<br />

h=0<br />

with the response propagator<br />

G 0 (q,ω)= [ −iω + Dq a (r + q 2 ) ] −1<br />

. (7.22)<br />

As is readily established by means of the residue theorem, its Fourier backtransform<br />

<strong>in</strong> time obeys causality,<br />

G 0 (q,t)=Θ(t)e −Dqa (r+q 2 ) t . (7.23)<br />

Sett<strong>in</strong>g h α = 0, and with the noise correlator (7.19) <strong>in</strong> Fourier space<br />

〈<br />

ζ α (q,ω) ζ β (q ′ ,ω ′ ) 〉 =2k B TDq a (2π) d+1 δ(q + q ′ ) δ(ω + ω ′ ) δ αβ , (7.24)


7 Field-Theory Approaches to Nonequilibrium Dynamics 301<br />

we obta<strong>in</strong> the Gaussian dynamic correlation function 〈 S α (q,ω) S β (q ′ ,ω ′ ) 〉 0 =<br />

C 0 (q,ω)(2π) d+1 δ(q + q ′ ) δ(ω + ω ′ ), where<br />

C 0 (q,ω)=<br />

2k B TDq a<br />

ω 2 +[Dq a (r + q 2 )] 2 =2k BTDq a |G 0 (q,ω)| 2 . (7.25)<br />

The fluctuation–dissipation theorem (7.13) is of course satisfied; moreover, as<br />

function of wavevector and time,<br />

C 0 (q,t)=<br />

k BT<br />

r + q 2 e−Dqa (r+q 2 ) |t| . (7.26)<br />

In the Gaussian approximation, away from criticality (r > 0, q ≠ 0) the<br />

temporal correlations for models A and B decay exponentially, with the relaxation<br />

rate ω c (r, q) = Dq 2+a (1 + r/q 2 ). Upon comparison with the dynamic<br />

scal<strong>in</strong>g hypothesis (7.9), we <strong>in</strong>fer the mean-field scal<strong>in</strong>g exponents<br />

ν 0 =1/2 andz 0 =2+a. At the critical po<strong>in</strong>t, a nonconserved order parameter<br />

relaxes diffusively (z 0 = 2) <strong>in</strong> this approximation, whereas the conserved<br />

order parameter k<strong>in</strong>etics becomes even slower, namely subdiffusive with<br />

z 0 = 4. F<strong>in</strong>ally, <strong>in</strong>vok<strong>in</strong>g Eqs. (7.12), (7.13), (7.14), or simply the static limit<br />

C 0 (q, 0) = k B T/(r + q 2 ), we f<strong>in</strong>d η 0 = 0 for the Gaussian model.<br />

The full nonl<strong>in</strong>ear Langev<strong>in</strong> equation (7.18) cannot be solved exactly. Yet<br />

a perturbation expansion with respect to the coupl<strong>in</strong>g u may be set up <strong>in</strong> a<br />

slightly cumbersome, but straightforward manner by direct iteration of the<br />

equations of motion [16, 17]. More elegantly, one may utilise a path-<strong>in</strong>tegral<br />

representation of the Langev<strong>in</strong> stochastic process [18, 19], which allows the<br />

application of all the standard tools from statistical and quantum field theory<br />

[1–6], and has the additional advantage of render<strong>in</strong>g symmetries <strong>in</strong> the<br />

problem more explicit [8, 9, 13].<br />

7.1.2 Field Theory Representation of Langev<strong>in</strong> Equations<br />

Our start<strong>in</strong>g po<strong>in</strong>t is a set of coupled Langev<strong>in</strong> equations of the form (7.15)<br />

for mesoscopic, coarse-gra<strong>in</strong>ed stochastic variables S α (x,t). For the stochastic<br />

forces, we make the simplest possible assumption of Gaussian white noise,<br />

〈ζ α (x,t)〉 =0,<br />

〈<br />

ζ α (x,t) ζ β (x ′ ,t ′ ) 〉 =2L α δ(x − x ′ ) δ(t − t ′ ) δ αβ , (7.27)<br />

where L α may represent a differential operator (such as the Laplacian ∇ 2 for<br />

conserved fields), and even a functional of S α . In the time <strong>in</strong>terval 0 ≤ t ≤ t f ,<br />

the moments (7.27) are encoded <strong>in</strong> the probability distribution<br />

[<br />

W[ζ] ∝ exp − 1 ∫ ∫ tf<br />

d d x dt ∑ ζ α (x,t) [ (L α ) −1 ζ α (x,t) ]] . (7.28)<br />

4 0 α<br />

If we now switch variables from the stochastic noise ζ α to the fields S α by<br />

means of the equations of motion (7.15), we obta<strong>in</strong>


302 U. C. Täuber<br />

W[ζ] D[ζ] =P[S] D[S] ∝ e −G[S] D[S] , (7.29)<br />

with the statistical weight determ<strong>in</strong>ed by the Onsager–Machlup functional [9]<br />

G[S] = 1 ∫ ∫<br />

d d x dt ∑ ( ( )]<br />

∂S<br />

α<br />

∂S<br />

− F α [S]<br />

)[(L α ) −1 α<br />

− F α [S] .<br />

4<br />

∂t<br />

∂t<br />

α<br />

(7.30)<br />

Note that the Jacobian for the nonl<strong>in</strong>ear variable transformation {ζ α }→<br />

{S α } has been omitted here. In fact, the above procedure is properly def<strong>in</strong>ed<br />

through appropriately discretis<strong>in</strong>g time. If a forward (Itô) discretisation is applied,<br />

then <strong>in</strong>deed the associated functional determ<strong>in</strong>ant is a mere constant<br />

that can be absorbed <strong>in</strong> the functional measure. The functional (7.30) already<br />

represents a desired field theory action. S<strong>in</strong>ce the probability distribution for<br />

the stochastic forces should be normalised, ∫ D[ζ] W [ζ] = 1, the associated<br />

“partition function” is unity, and carries no physical <strong>in</strong>formation (as opposed<br />

to static statistical field theory, where it determ<strong>in</strong>es the free energy and hence<br />

the entire thermodynamics). The Onsager–Machlup representation is however<br />

plagued by technical problems: Eq. (7.30) conta<strong>in</strong>s (L α ) −1 , which for<br />

conserved variables entails the <strong>in</strong>verse Laplacian operator, i.e., a Green function<br />

<strong>in</strong> real space or the s<strong>in</strong>gular factor 1/q 2 <strong>in</strong> Fourier space; moreover the<br />

nonl<strong>in</strong>earities <strong>in</strong> F α [S] appear quadratically. Hence it is desirable to l<strong>in</strong>earise<br />

the action (7.30) by means of a Hubbard–Stratonovich transformation [9].<br />

We shall follow an alternative, more general route that completely avoids<br />

the appearance of the <strong>in</strong>verse operators (L α ) −1 <strong>in</strong> <strong>in</strong>termediate steps. Our<br />

goal is to average over noise “histories” for observables A[S] that need to be<br />

expressible <strong>in</strong> terms of the stochastic fields S α : 〈A[S]〉 ζ ∝ ∫ D[ζ] A[S(ζ)] W [ζ].<br />

For this purpose, we employ the identity<br />

∫<br />

1= D[S] ∏ ∏<br />

( ∂S α )<br />

(x,t)<br />

δ<br />

− F α [S](x,t) − ζ α (x,t)<br />

∂t<br />

α (x,t)<br />

∫ ∫ [ ∫ ∫<br />

= D[i ˜S] D[S]exp − d d x dt ∑ ( )]<br />

∂S<br />

˜S α α<br />

− F α [S] − ζ α , (7.31)<br />

∂t<br />

α<br />

where the first l<strong>in</strong>e constitutes a rather <strong>in</strong>volved representation of the unity (<strong>in</strong><br />

a somewhat symbolic notation; aga<strong>in</strong> proper discretisation should be <strong>in</strong>voked<br />

here), and the second l<strong>in</strong>e utilises the Fourier representation of the (functional)<br />

delta distribution by means of the purely imag<strong>in</strong>ary auxiliary fields ˜S (and<br />

factors 2π have been absorbed <strong>in</strong> its functional measure).<br />

Insert<strong>in</strong>g (7.31) and the probability distribution (7.28) <strong>in</strong>to the desired<br />

stochastic noise average, we arrive at<br />

∫ ∫ [ ∫ ∫<br />

〈A[S]〉 ζ ∝ D[i ˜S] D[S]exp − d d x dt ∑ ( )]<br />

∂S<br />

˜S α α<br />

− F α [S] A[S]<br />

∂t<br />

α<br />

∫ ( ∫ ∫<br />

× D[ζ]exp − d d x dt ∑ [ ])<br />

1<br />

4 ζα (L α ) −1 ζ α − ˜S α ζ α . (7.32)<br />

α


7 Field-Theory Approaches to Nonequilibrium Dynamics 303<br />

We may now evaluate the Gaussian <strong>in</strong>tegrals over the noise ζ α , which yields<br />

∫<br />

∫<br />

〈A[S]〉 ζ = D[S] A[S] P[S] , P[S] ∝ D[i ˜S]e −A[ ˜S,S] , (7.33)<br />

with the statistical weight now governed by the Janssen–De Dom<strong>in</strong>icis “response”<br />

functional [9, 18, 19]<br />

∫ ∫<br />

A[ ˜S,S]=<br />

tf<br />

d d x dt ∑ [ ( )<br />

]<br />

∂S<br />

˜S α α<br />

− F α [S] −<br />

0<br />

∂t<br />

˜S α L α ˜Sα . (7.34)<br />

α<br />

Once aga<strong>in</strong>, we have omitted the functional determ<strong>in</strong>ant from the variable<br />

change {ζ α }→{S α }, and normalisation implies ∫ D[i ˜S] ∫ D[S]e −A[ ˜S,S] =1.<br />

The first term <strong>in</strong> the action (7.34) encodes the temporal evolution accord<strong>in</strong>g<br />

to the systematic terms <strong>in</strong> the Langev<strong>in</strong> equations (7.15), whereas the second<br />

term specifies the noise correlations (7.27). S<strong>in</strong>ce the auxiliary variables ˜S α ,<br />

often termed Mart<strong>in</strong>–Siggia–Rose response fields [20], appear only quadratically<br />

here, they may be elim<strong>in</strong>ated via complet<strong>in</strong>g the squares and Gaussian<br />

<strong>in</strong>tegrations; thereby one recovers the Onsager–Machlup functional (7.30).<br />

The Janssen–De Dom<strong>in</strong>icis functional (7.34) takes the form of a (d + 1)-<br />

dimensional statistical field theory with two <strong>in</strong>dependent sets of fields S α and<br />

˜S α . We may thus br<strong>in</strong>g the established mach<strong>in</strong>ery of statistical and quantum<br />

field theory [1–6] to bear here; it should however be noted that the response<br />

functional formalism for stochastic Langev<strong>in</strong> dynamics <strong>in</strong>corporates causality<br />

<strong>in</strong> a nontrivial manner, which leads to important dist<strong>in</strong>ctions [8].<br />

Let us specify the Janssen–De Dom<strong>in</strong>icis functional for the purely relaxational<br />

models A and B [16, 17], see Eqs. (7.18) and (7.19), splitt<strong>in</strong>g it <strong>in</strong>to<br />

the Gaussian and anharmonic parts A = A 0 + A <strong>in</strong>t [9], which read<br />

∫ ∫<br />

A 0 [ ˜S,S] = d d x dt ∑ ( [ ]<br />

∂<br />

˜S α ∂t + D (i∇)a (r − ∇ 2 ) S α<br />

α<br />

)<br />

−D ˜S α (i∇) a ˜Sα − D ˜S α (i∇) a h α , (7.35)<br />

A <strong>in</strong>t [ ˜S,S] =D u ∫ ∫<br />

d d x dt ∑ ˜S α (i∇) a S α S β S β . (7.36)<br />

6<br />

α,β<br />

S<strong>in</strong>ce we are <strong>in</strong>terested <strong>in</strong> the vic<strong>in</strong>ity of the critical po<strong>in</strong>t T ≈ T c ,wehave<br />

absorbed the constant k B T c <strong>in</strong>to the fields. The prescription (7.33) tells us how<br />

to compute time-dependent correlation functions 〈 S α (x,t) S β (x ′ ,t ′ ) 〉 .Us<strong>in</strong>g<br />

Eq. (7.35), the dynamic order parameter susceptibility follows from<br />

χ αβ (x − x ′ ,t− t ′ )= δ〈Sα (x,t)〉<br />

δh β (x ′ ,t ′ )<br />

∣ = D<br />

h=0<br />

〈<br />

〉<br />

S α (x,t)(i∇) a ˜Sβ (x ′ ,t ′ )<br />

(7.37)<br />

for the simple relaxational models (only), the response function is just given by<br />

a correlator that <strong>in</strong>volves an auxiliary variable, which expla<strong>in</strong>s why the ˜S α are<br />

;


304 U. C. Täuber<br />

referred to as “response” fields. In equilibrium, one may employ the Onsager–<br />

Machlup functional (7.30) to derive the fluctuation–dissipation theorem [9]<br />

χ αβ (x − x ′ ,t− t ′ )=Θ(t − t ′ ) ∂<br />

∂t ′ 〈<br />

S α (x,t) S β (x ′ ,t ′ ) 〉 , (7.38)<br />

which is equivalent to Eq. (7.13) <strong>in</strong> Fourier space.<br />

In order to access arbitrary correlators, we def<strong>in</strong>e the generat<strong>in</strong>g functional<br />

〈 ∫ ∫<br />

Z[˜j,j]= exp d d x dt ∑ (˜j α ˜Sα + j α S α)〉 , (7.39)<br />

α<br />

wherefrom the correlation functions follow via functional derivatives,<br />

〈 ∏<br />

〉<br />

S αi ˜S αj = ∏ δ δ<br />

Z[˜j,j] ∣ , (7.40)<br />

δj αi δ˜j<br />

ij<br />

ij<br />

αj ∣˜j=0=j<br />

and the cumulants or connected correlation functions via<br />

〈 ∏<br />

〉<br />

S αi ˜S αj = ∏ δ δ<br />

ln Z[˜j,j] ∣<br />

ij<br />

c<br />

δj αi δ˜j<br />

ij<br />

αj<br />

∣˜j=0=j<br />

. (7.41)<br />

In the harmonic approximation, sett<strong>in</strong>g u =0,Z[˜j,j] can be evaluated explicitly<br />

(most directly <strong>in</strong> Fourier space) by means of Gaussian <strong>in</strong>tegration [9,13];<br />

one thereby recovers (with k B T = 1) the Gaussian response propagator (7.22)<br />

and two-po<strong>in</strong>t 〈 correlation function (7.25). Moreover, as a consequence of<br />

causality, ˜Sα (q,ω) ˜S β (q ′ ,ω )〉<br />

′ =0.<br />

7.1.3 Outl<strong>in</strong>e of Dynamic Perturbation Theory<br />

0<br />

S<strong>in</strong>ce we cannot evaluate correlation functions with the nonl<strong>in</strong>ear action (7.36)<br />

exactly, we resort to a perturbational treatment, assum<strong>in</strong>g, for the time be<strong>in</strong>g,<br />

a small coupl<strong>in</strong>g strength u. Theperturbation expansion with respect to u is<br />

constructed by rewrit<strong>in</strong>g the desired correlation functions <strong>in</strong> terms of averages<br />

with respect to the Gaussian action (7.35), henceforth <strong>in</strong>dicated with <strong>in</strong>dex<br />

’0’, and then expand<strong>in</strong>g the exponential of −A <strong>in</strong>t ,<br />

〈 ∏<br />

ij<br />

S αi ˜S αj 〉<br />

=<br />

〈 ∏ ˜S ij Sαi αj −A<strong>in</strong>t[ ˜S,S]〉<br />

e<br />

〈 ∏<br />

=<br />

ij<br />

〈 〉<br />

e−A<strong>in</strong>t[ ˜S,S]<br />

0<br />

S αi ˜S ∑ ∞ αj<br />

l=0<br />

1<br />

l!<br />

0<br />

(<br />

−A <strong>in</strong>t [ ˜S,S]<br />

) l<br />

〉<br />

0<br />

. (7.42)<br />

The rema<strong>in</strong><strong>in</strong>g Gaussian averages, a series of polynomials <strong>in</strong> the fields S α<br />

and ˜S α , can be evaluated by means of Wick’s theorem, here an immediate


7 Field-Theory Approaches to Nonequilibrium Dynamics 305<br />

consequence of the Gaussian statistical weight, which states that all such averages<br />

can be written as a sum over all possible factorisations <strong>in</strong>to Gaussian<br />

two-po<strong>in</strong>t functions 〈S α ˜Sβ 〉 0 , i.e., essentially the response propagator G 0 ,<br />

Eq. (7.22), and 〈S α S β 〉 0 , the Gaussian correlation function C 0 , Eq. (7.25).<br />

Recall that the denom<strong>in</strong>ator <strong>in</strong> Eq. (7.42) is exactly unity as a consequence of<br />

normalisation; alternatively, this result follows from causality <strong>in</strong> conjunction<br />

with our forward descretisation prescription, which implies that we should<br />

identify Θ(0) = 0. (We remark that had we chosen another temporal discretisation<br />

rule, any apparent contributions from the denom<strong>in</strong>ator would be<br />

precisely cancelled by the <strong>in</strong> this case nonvanish<strong>in</strong>g functional Jacobian from<br />

the variable transformation {ζ α }→{S α }.) At any rate, our stochastic field<br />

theory conta<strong>in</strong>s no “vacuum” contributions.<br />

The many terms <strong>in</strong> the perturbation expansion (7.42) are most lucidly<br />

organised <strong>in</strong> a graphical representation, us<strong>in</strong>g Feynman diagrams with the<br />

basic elements depicted <strong>in</strong> Fig. 7.1. We represent the response propagator<br />

(7.22) by a directed l<strong>in</strong>e (here conventionally from right to left), which encodes<br />

its causal nature; the noise by a two-po<strong>in</strong>t “source” vertex, and the anharmonic<br />

term <strong>in</strong> Eq. (7.36) as a four-po<strong>in</strong>t vertex. In the diagrams represent<strong>in</strong>g the<br />

different terms <strong>in</strong> the perturbation series, these vertices serve as l<strong>in</strong>ks for the<br />

propagator l<strong>in</strong>es, with the fields S α be<strong>in</strong>g encoded as the “<strong>in</strong>com<strong>in</strong>g”, and the<br />

˜S α as the “outgo<strong>in</strong>g” components of the l<strong>in</strong>es. In Fourier space, translational<br />

<strong>in</strong>variance <strong>in</strong> space and time implies wavevector and frequency conservation<br />

at each vertex, see Fig. 7.2 below. An alternative, equivalent representation<br />

uses both the response and correlation propagators as <strong>in</strong>dependent elements,<br />

the latter depicted as undirected l<strong>in</strong>e, thereby dispos<strong>in</strong>g of the noise vertex,<br />

and reta<strong>in</strong><strong>in</strong>g the nonl<strong>in</strong>earity <strong>in</strong> Fig. 7.1(c) as sole vertex.<br />

Follow<strong>in</strong>g standard field theory procedures [1–5], one establishes that the<br />

perturbation series for the cumulants (7.41) is given <strong>in</strong> terms of connected<br />

Feynman graphs only (for a detailed exposition of this and the follow<strong>in</strong>g results,<br />

see Ref. [13]). An additional helpful reduction <strong>in</strong> the number of diagrams<br />

to be considered arises when one considers the vertex functions, which generalise<br />

the self-energy contributions Σ(q,ω)<strong>in</strong>theDyson equation for the<br />

response propagator, G(q,ω) −1 = Dq a χ(q,ω) −1 = G 0 (q,ω) −1 − Σ(q,ω).<br />

(a)<br />

α<br />

q,ω<br />

β<br />

=<br />

1<br />

i ω+ D q<br />

a<br />

( r + q2)<br />

δ αβ<br />

(b)<br />

α<br />

β<br />

q<br />

-q<br />

=<br />

2 D q a δ αβ<br />

(c)<br />

α<br />

q<br />

α<br />

β<br />

β<br />

=<br />

D q a<br />

u<br />

6<br />

Fig. 7.1. Elements of dynamic perturbation theory for the O(n)-symmetric relaxational<br />

models: (a) response propagator; (b) noisevertex;(c) anharmonic vertex


306 U. C. Täuber<br />

To this end, we def<strong>in</strong>e the fields ˜Φ α = δ ln Z/δ˜j α and Φ α = δ ln Z/δj α ,and<br />

<strong>in</strong>troduce the new generat<strong>in</strong>g functional<br />

∫ ∫<br />

Γ [ ˜Φ, Φ] =− ln Z[˜j,j]+ d d x dt ∑ (˜j α ˜Φα + j α Φ α) , (7.43)<br />

α<br />

wherefrom the vertex functions are obta<strong>in</strong>ed via the functional derivatives<br />

Γ (Ñ,N)<br />

{α i};{α j} = Ñ<br />

∏<br />

i<br />

N<br />

δ<br />

αi<br />

δ ˜Φ<br />

∏<br />

δ<br />

δΦ Γ [ ˜Φ, Φ] ∣<br />

αj<br />

j<br />

∣˜j=0=j<br />

. (7.44)<br />

Diagrammatically, these quantities turn out to be represented by the possible<br />

sets of one-particle (1PI) irreducible Feynman graphs with N <strong>in</strong>com<strong>in</strong>g and<br />

Ñ outgo<strong>in</strong>g “amputated” legs; i.e., these diagrams do not split <strong>in</strong>to allowed<br />

subgraphs by simply cutt<strong>in</strong>g any s<strong>in</strong>gle propagator l<strong>in</strong>e. For example, for the<br />

two-po<strong>in</strong>t functions a direct calculation yields the relations<br />

Γ (1,1) (q,ω)=Dq a χ(−q, −ω) −1 = G 0 (−q, −ω) −1 − Σ(−q, −ω) ,(7.45)<br />

Γ (2,0) (q,ω)=−<br />

C(q,ω) qa<br />

= −2D Im Γ (1,1) (q,ω) , (7.46)<br />

|G(q,ω)|<br />

2<br />

ω<br />

where the second equation for Γ (2,0) follows from the fluctuation–dissipation<br />

theorem (7.13). Note that Γ (0,2) (q,ω) = 0 vanishes because of causality.<br />

The perturbation series can then be organised graphically as an expansion<br />

<strong>in</strong> successive orders with respect to the number of closed propagator loops.<br />

As an example, Fig. 7.2 depicts the one-loop contributions for the vertex<br />

functions Γ (1,1) and Γ (1,3) <strong>in</strong> the time doma<strong>in</strong> with all required labels. One<br />

may formulate general Feynman rules for the construction of the diagrams and<br />

their translation <strong>in</strong>to mathematical expressions for the lth order contribution<br />

to the vertex function Γ (Ñ,N) :<br />

1. Draw all topologically different, connected one-particle irreducible graphs<br />

with Ñ outgo<strong>in</strong>g and N <strong>in</strong>com<strong>in</strong>g l<strong>in</strong>es connect<strong>in</strong>g l relaxation vertices ∝<br />

u. Donot allow closed response loops (s<strong>in</strong>ce <strong>in</strong> the Itô calculus Θ(0) = 0).<br />

2. Attach wavevectors q i , frequencies ω i or times t i , and component <strong>in</strong>dices<br />

α i to all directed l<strong>in</strong>es, obey<strong>in</strong>g “momentum (and energy)” conservation<br />

at each vertex.<br />

(a)<br />

α<br />

k<br />

t<br />

t´<br />

β -k<br />

α<br />

(b)<br />

α<br />

0<br />

q-k<br />

β<br />

t<br />

k<br />

α<br />

α<br />

α<br />

-k<br />

t´<br />

Fig. 7.2. One-loop diagrams for (a) Γ (1,1) and (b) Γ (1,3) <strong>in</strong> the time doma<strong>in</strong>


7 Field-Theory Approaches to Nonequilibrium Dynamics 307<br />

3. Each directed l<strong>in</strong>e corresponds to a response propagator G 0 (−q, −ω) or<br />

G 0 (q,t i −t j ) <strong>in</strong> the frequency and time doma<strong>in</strong>, respectively, the two-po<strong>in</strong>t<br />

vertex to the noise strength 2D q a , and the four-po<strong>in</strong>t relaxation vertex<br />

to −D q a u/6. Closed loops imply <strong>in</strong>tegrals over the <strong>in</strong>ternal wavevectors<br />

and frequencies or times, subject to causality constra<strong>in</strong>ts, as well as sums<br />

over the <strong>in</strong>ternal vector <strong>in</strong>dices. Apply the residue theorem to evaluate<br />

frequency <strong>in</strong>tegrals.<br />

4. Multiply with −1 and the comb<strong>in</strong>atorial factor count<strong>in</strong>g all possible ways<br />

of connect<strong>in</strong>g the propagators, l relaxation vertices, and k two-po<strong>in</strong>t vertices<br />

lead<strong>in</strong>g to topologically identical graphs, <strong>in</strong>clud<strong>in</strong>g a factor 1/l! k!<br />

orig<strong>in</strong>at<strong>in</strong>g <strong>in</strong> the expansion of exp(−A <strong>in</strong>t [ ˜S,S]).<br />

For later use, we provide the explicit results for the two-po<strong>in</strong>t vertex functions<br />

to two-loop order. After some algebra, the three diagrams <strong>in</strong> Fig. 7.3<br />

give<br />

[<br />

Γ (1,1) (q,ω)=iω + Dq a r + q 2 + n +2<br />

6<br />

( ) 2 ∫<br />

n +2<br />

− u<br />

6<br />

k<br />

− n +2<br />

×<br />

18 u2 ∫k<br />

(<br />

1 −<br />

∫<br />

1<br />

r + k 2<br />

∫<br />

1<br />

r + k 2<br />

∫<br />

u<br />

k<br />

k ′ 1<br />

1<br />

r + k 2<br />

1<br />

k ′ (r + k ′2 ) 2<br />

1<br />

r + k ′ 2<br />

r +(q − k − k ′ )<br />

)]<br />

2<br />

, (7.47)<br />

iω<br />

iω + ∆(k)+∆(k ′ )+∆(q − k − k ′ )<br />

where we have separated out the dynamic part <strong>in</strong> the last l<strong>in</strong>e, and <strong>in</strong>troduced<br />

the abbreviations ∆(q) =Dq a (r + q 2 )and ∫ k = ∫ d d k/(2π) d [13]. For the<br />

noise vertex, Fig. 7.4(a) yields [13]<br />

[<br />

Γ (2,0) (q,ω)=−2Dq a 1+Dq a n +2<br />

18<br />

∫k<br />

u2<br />

1<br />

×<br />

r +(q − k − k ′ ) 2 Re 1<br />

∫<br />

1<br />

r + k 2<br />

k ′ 1<br />

r + k ′ 2<br />

]<br />

iω + ∆(k)+∆(k ′ )+∆(q − k − k ′ ; (7.48)<br />

)<br />

notice that for model B, as a consequence of the conservation law for the<br />

order parameter and ensu<strong>in</strong>g wavevector dependence of the nonl<strong>in</strong>ear vertex,<br />

+ + + ...<br />

Fig. 7.3. One-particle irreducible diagrams for Γ (1,1) (q,ω) to second order <strong>in</strong> u


308 U. C. Täuber<br />

(a)<br />

(b)<br />

Fig. 7.4. (a) Two-loop diagram for Γ (2,0) (q,ω); (b) one-loop graph for Γ (1,3)<br />

see Fig. 7.1(c), to all orders <strong>in</strong> the perturbation expansion<br />

a =2 : Γ (1,1) ∂<br />

(q =0,ω)=iω,<br />

∂q 2 Γ (2,0) (q,ω)<br />

∣ = −2D . (7.49)<br />

q=0<br />

At last, with the shorthand notation k =(q,ω), the analytical expression<br />

correspond<strong>in</strong>g to the graph <strong>in</strong> Fig. 7.4(b) for the four-po<strong>in</strong>t vertex function<br />

at symmetrically chosen external wavevector labels reads<br />

( ) a [ 3<br />

Γ (1,3) (−3k/2; {k/2}) =D<br />

2 q u<br />

∫<br />

(<br />

1 1<br />

×<br />

k r + k 2 r +(q − k) 2 1 −<br />

7.1.4 Renormalisation<br />

1 − n +8<br />

6<br />

u<br />

iω<br />

iω + ∆(k)+∆(q − k)<br />

)]<br />

. (7.50)<br />

Consider a typical loop <strong>in</strong>tegral, say the correction <strong>in</strong> Eq. (7.50) to the fourpo<strong>in</strong>t<br />

vertex function Γ (1,3) at zero external frequency and momentum, whose<br />

“bare” value, without any fluctuation contributions, is u. In dimensions d


7 Field-Theory Approaches to Nonequilibrium Dynamics 309<br />

Conversely, for dimensions d ≥ 4, the <strong>in</strong>tegral <strong>in</strong> (7.51) develops ultraviolet<br />

(UV) divergences as the upper <strong>in</strong>tegral boundary is sent to <strong>in</strong>f<strong>in</strong>ity (k = |k|),<br />

∫ Λ<br />

k d−1 { }<br />

ln(Λ<br />

(r + k 2 ) 2 dk ∼ 2 /r) d =4<br />

Λ d−4 →∞ as Λ →∞. (7.52)<br />

d>4<br />

0<br />

In lattice models, there is a f<strong>in</strong>ite wavevector cutoff, namely the Brillou<strong>in</strong><br />

zone boundary, Λ ∼ (2π/a 0 ) d for a hypercubic lattice with lattice constant<br />

a 0 , whence physically these UV problems do not emerge. Yet we shall see that<br />

a formal treatment of these unphysical UV divergences will allow us to <strong>in</strong>fer<br />

the correct power laws for the physical IR s<strong>in</strong>gularities associated with the<br />

critical po<strong>in</strong>t. The borderl<strong>in</strong>e dimension that separates the IR and UV s<strong>in</strong>gular<br />

regimes is referred to as upper critical dimension d c ; here d c = 4. Note that<br />

at d c , UV and IR s<strong>in</strong>gularities are <strong>in</strong>timately connected and appear <strong>in</strong> the<br />

form of logarithmic divergences, see Eq. (7.52). The situation is summarised<br />

<strong>in</strong> Table 7.1, where we have also stated that models with cont<strong>in</strong>uous order<br />

parameter symmetry, such as the Hamiltonian (7.7) with n ≥ 2, do not allow<br />

long-range order <strong>in</strong> dimensions d ≤ d lc = 2 (Merm<strong>in</strong>–Wagner–Hohenberg<br />

theorem [21–23]). Here, d lc is called the lower critical dimension; for the Is<strong>in</strong>g<br />

model represented by Eq. (7.7) with n =1,ofcoursed lc =1.<br />

Table 7.1. Mathematical and physical dist<strong>in</strong>ctions of the regimes dd c, for the O(n)-symmetric models A and B (or static Φ 4 field theory)<br />

Dimension Perturbation Model A / B or Critical<br />

Interval Series Φ 4 Field Theory Behaviour<br />

d ≤ d lc = 2 IR-s<strong>in</strong>gular ill-def<strong>in</strong>ed no long-range<br />

UV-convergent u relevant order (n ≥ 2)<br />

2


310 U. C. Täuber<br />

phase transition; but u becomes irrelevant for d>4: one then expects meanfield<br />

(Gaussian) critical exponents. At the upper critical dimension d c =4,<br />

the nonl<strong>in</strong>ear coupl<strong>in</strong>g u is marg<strong>in</strong>ally relevant: this will <strong>in</strong>duce logarithmic<br />

corrections to the mean-field scal<strong>in</strong>g laws, see Table 7.1.<br />

It is obviously not a simple task to treat the IR-s<strong>in</strong>gular perturbation<br />

expansion <strong>in</strong> a mean<strong>in</strong>gful, well-def<strong>in</strong>ed manner, and thus allow nonanalytic<br />

modifications of the critical power laws (note that mean-field scal<strong>in</strong>g is completely<br />

determ<strong>in</strong>ed by dimensional analysis or power count<strong>in</strong>g). The key of<br />

the success of the RG approach is to focus on the very specific symmetry<br />

that emerges near critical po<strong>in</strong>ts, namely scale <strong>in</strong>variance. There are several<br />

(largely equivalent) versions of the RG method; we shall here formulate<br />

and employ the field-theoretic variant [1–6, 8, 13]. In order to proceed,<br />

it is convenient to evaluate the loop <strong>in</strong>tegrals <strong>in</strong> momentum space by means<br />

of dimensional regularisation, whereby one assigns f<strong>in</strong>ite values even to UVdivergent<br />

expressions, namely the analytically cont<strong>in</strong>ued values from the UVf<strong>in</strong>ite<br />

range. For example, even for non<strong>in</strong>teger dimensions d and σ, weset<br />

∫<br />

d d k<br />

(2π) d<br />

k 2σ Γ (σ + d/2) Γ (s − σ − d/2)<br />

(τ + k 2 ) s =<br />

2 d π d/2 Γ (d/2) Γ (s)<br />

The renormalisation program then consists of the follow<strong>in</strong>g steps:<br />

τ σ−s+d/2 . (7.53)<br />

1. We aim to carefully keep track of formal, unphysical UV divergences. In<br />

dimensionally regularised <strong>in</strong>tegrals (7.53), these appear as poles <strong>in</strong> ɛ =<br />

d c − d; their residues characterise the asymptotic UV behaviour of the field<br />

theory under consideration.<br />

2. Therefrom we may <strong>in</strong>fer the (UV) scal<strong>in</strong>g properties of the control parameters<br />

of the model under a RG transformation, namely essentially a change<br />

of the momentum scale µ, while keep<strong>in</strong>g the form of the action <strong>in</strong>variant.<br />

This will allow us to def<strong>in</strong>e suitable runn<strong>in</strong>g coupl<strong>in</strong>gs.<br />

3. We seek fixed po<strong>in</strong>ts <strong>in</strong> parameter space where certa<strong>in</strong> marg<strong>in</strong>al coupl<strong>in</strong>gs<br />

(u here) do not change anymore under RG transformations. This describes<br />

a scale-<strong>in</strong>variant regime for the model under consideration, where the UV<br />

and IR scal<strong>in</strong>g properties become <strong>in</strong>timately l<strong>in</strong>ked. Study<strong>in</strong>g the parameter<br />

flows near a stable RG fixed po<strong>in</strong>t then allows us to extract the<br />

asymptotic IR power laws.<br />

As a prelim<strong>in</strong>ary step, we need to take <strong>in</strong>to account that the fluctuations<br />

will also shift the critical po<strong>in</strong>t downwards from the mean-field phase transition<br />

temperature T 0 c ; i.e., we expect the transition to occur at T c


7 Field-Theory Approaches to Nonequilibrium Dynamics 311<br />

Notice that this quantity depends on microscopic details (the lattice structure<br />

enters the cutoff Λ) and is thus not universal; moreover it diverges for d ≥ 2<br />

(quadratically near d c =4)asΛ →∞.Wenextuser = τ + r c to write<br />

physical quantities as functions of the true distance τ from the critical po<strong>in</strong>t,<br />

which technically amounts to an additive renormalisation; e.g., the dynamic<br />

response function becomes to one-loop order<br />

χ(q,ω) −1 = − iω<br />

Dq a + q2 + τ<br />

[<br />

1 − n +2 u<br />

6<br />

∫<br />

k<br />

]<br />

1<br />

k 2 (τ + k 2 + O(u 2 ) . (7.55)<br />

)<br />

The rema<strong>in</strong><strong>in</strong>g loop <strong>in</strong>tegral is UV-s<strong>in</strong>gular <strong>in</strong> dimensions d ≥ d c =4.<br />

We may now formally absorb the rema<strong>in</strong><strong>in</strong>g UV divergences <strong>in</strong>to renormalised<br />

fields and parameters, a procedure called multiplicative renormalisation.<br />

For the renormalised fields, we use the convention<br />

SR α = Z 1/2<br />

S<br />

S α , ˜Sα R = Z 1/2 ˜S α , (7.56)<br />

where we have exploited the O(n) rotational symmetry <strong>in</strong> us<strong>in</strong>g identical<br />

renormalisation constants (Z factors) for each component. The renormalised<br />

cumulants with N order parameter fields S α and Ñ response fields ˜S α naturally<br />

<strong>in</strong>volve the product Z N/2<br />

S<br />

Γ (Ñ,N)<br />

R<br />

ZÑ/2 , whence<br />

˜S<br />

˜S<br />

= Z −Ñ/2 Z −N/2<br />

˜S S<br />

Γ (Ñ,N) . (7.57)<br />

In a similar manner, we relate the “bare” parameters of the theory via Z<br />

factors to their renormalised counterparts, which we furthermore render dimensionless<br />

through appropriate momentum scale factors,<br />

D R = Z D D, τ R = Z τ τµ −2 , u R = Z u uA d µ d−4 , (7.58)<br />

where we have separated out the factor A d = Γ (3 − d/2)/2 d−1 π d/2 for convenience.<br />

In the m<strong>in</strong>imal subtraction scheme, the Z factors conta<strong>in</strong> only the<br />

UV-s<strong>in</strong>gular terms, which <strong>in</strong> dimensional regularisation appear as poles at<br />

ɛ = 0, and their residues, evaluated at d = d c .<br />

These renormalisation constants are not all <strong>in</strong>dependent, however; s<strong>in</strong>ce<br />

the equilibrium fluctuation–dissipation theorem (7.38) or (7.13) must hold <strong>in</strong><br />

the renormalised theory as well, we <strong>in</strong>fer that necessarily<br />

and consequently from Eq. (7.45)<br />

Z D = ( Z S /Z ˜S) 1/2<br />

, (7.59)<br />

χ R = Z S χ. (7.60)<br />

Moreover, for model B with conserved order parameter Eq. (7.49) implies that<br />

to all orders <strong>in</strong> the perturbation expansion


312 U. C. Täuber<br />

a =2 : Z ˜S<br />

Z S =1, Z D = Z S . (7.61)<br />

For the follow<strong>in</strong>g, it is crucial that the theory is renormalisable, i.e., a f<strong>in</strong>ite<br />

number of reparametrisations suffice to formally rid it of all UV divergences.<br />

Indeed, for the relaxational models A and B, and the static G<strong>in</strong>zburg–Landau–<br />

Wilson Hamiltonian (7.7), all higher vertex function beyond the four-po<strong>in</strong>t<br />

function are UV-convergent near d c , and there are only the three <strong>in</strong>dependent<br />

static renormalisation factors Z S , Z τ ,andZ u , and <strong>in</strong> addition Z D for nonconserved<br />

order parameter dynamics. As we shall see, these directly translate<br />

<strong>in</strong>to the two <strong>in</strong>dependent static critical exponents and the unrelated dynamic<br />

scal<strong>in</strong>g exponent z for model A; for model B with conserved order parameter,<br />

Eq. (7.61) will yield a scal<strong>in</strong>g relation between z and η.<br />

In order to explicitly determ<strong>in</strong>e the renormalisation constants, we need<br />

to ensure that we stay away from the IR-s<strong>in</strong>gular regime. This is guaranteed<br />

by select<strong>in</strong>g as normalisation po<strong>in</strong>t either τ R = 1 (i.e., Z τ τ = µ 2 )orq = µ.<br />

Inevitably therefore, the renormalised theory depends on the correspond<strong>in</strong>g<br />

arbitrary momentum scale µ. S<strong>in</strong>ce there are no fluctuation contributions to<br />

order u to either ∂Γ (1,1) (q, 0)/∂q 2 or ∂Γ (1,1) (0,ω)/∂ω (at τ R = 1), we f<strong>in</strong>d<br />

Z S =1andZ D = 1 with<strong>in</strong> the one-loop approximation. Expressions (7.55)<br />

and (7.50) then yield with the formula (7.53)<br />

Z τ =1− n +2<br />

6<br />

u R<br />

ɛ<br />

, Z u =1− n +8<br />

6<br />

u R<br />

ɛ<br />

. (7.62)<br />

To two-loop order, we may <strong>in</strong>fer the field renormalisation Z S from the static<br />

susceptibility as the s<strong>in</strong>gular contributions to ∂χ R (q, 0)/∂q 2 | q=0 ,andZ D for<br />

model A, through a somewhat lengthy calculation [13], from either Γ (2,0)<br />

R<br />

(0, 0)<br />

or Γ (1,1)<br />

R<br />

(0,ω), with the results<br />

Z S =1+ n +2<br />

144<br />

u 2 R<br />

ɛ<br />

, a =0 : Z D =1− n +2<br />

144<br />

7.1.5 Scal<strong>in</strong>g Laws and Critical Exponents<br />

(<br />

6ln 4 3 − 1 ) u<br />

2<br />

R<br />

ɛ<br />

. (7.63)<br />

We now wish to related the renormalised vertex functions at different <strong>in</strong>verse<br />

length scales µ. This is accomplished by simply recall<strong>in</strong>g that the unrenormalised<br />

vertex functions obviously do not depend on µ,<br />

0=µ d<br />

dµ Γ (Ñ,N) (D, τ, u) =µ d [<br />

dµ<br />

ZÑ/2 ˜S<br />

]<br />

Z N/2<br />

S<br />

Γ (Ñ,N)<br />

R<br />

(µ, D R ,τ R ,u R ) . (7.64)<br />

In the second step, the bare quantities have been replaced with their renormalised<br />

counterparts. The <strong>in</strong>nocuous statement (7.64) then implies a very<br />

nontrivial partial differential equation for the renormalised vertex functions,<br />

the desired renormalisation group equation,


7 Field-Theory Approaches to Nonequilibrium Dynamics 313<br />

[<br />

µ ∂<br />

∂µ + Ñγ˜S + Nγ S<br />

∂<br />

+ γ D D R + γ τ τ R<br />

2<br />

∂D R<br />

∂<br />

∂τ R<br />

+ β u<br />

]<br />

∂<br />

∂u R<br />

× Γ (Ñ,N)<br />

R<br />

(µ, D R ,τ R ,u R )=0. (7.65)<br />

Here we have def<strong>in</strong>ed Wilson’s flow functions (the <strong>in</strong>dex “0” <strong>in</strong>dicates that<br />

the derivatives with respect to µ are to be taken with fixed unrenormalised<br />

parameters)<br />

γ ˜S<br />

= µ ∂<br />

∂µ ∣ ln Z ˜S<br />

, γ S = µ ∂<br />

0<br />

∂µ ∣ ln Z S , (7.66)<br />

0<br />

γ τ = µ ∂<br />

∂µ ∣ ln(τ R /τ) =−2+µ ∂<br />

0<br />

∂µ ∣ ln Z τ , (7.67)<br />

0<br />

γ D = µ ∂<br />

∂µ ∣ ln(D R /D) = 1 ( )<br />

γS − γ<br />

0<br />

2<br />

˜S , (7.68)<br />

where we have used the relation (7.59); for model B, Eq. (7.61) gives <strong>in</strong> addition<br />

γ D = γ S = −γ ˜S<br />

. (7.69)<br />

We have also <strong>in</strong>troduced the RG beta function for the nonl<strong>in</strong>ear coupl<strong>in</strong>g u,<br />

β u = µ<br />

∂<br />

(<br />

∂µ ∣ u R = u R d − 4+µ<br />

∂<br />

)<br />

0<br />

∂µ ∣ ln Z u . (7.70)<br />

0<br />

Explicitly, Eqs. (7.63) and (7.62) yield to lowest nontrivial order, with ɛ =<br />

4 − d,<br />

γ S = − n +2<br />

72 u2 R + O(u 3 R) , (7.71)<br />

a =0 : γ D = n +2 (6ln 4 )<br />

72 3 − 1 u 2 R + O(u 3 R) , (7.72)<br />

γ τ = −2+ n +2 u R + O(u 2<br />

6<br />

R) , (7.73)<br />

[<br />

β u = u R −ɛ + n +8<br />

]<br />

u R + O(u 2<br />

6<br />

R) . (7.74)<br />

In the RG equation for the renormalised dynamic susceptibility, Eq. (7.60)<br />

tells us that the second term <strong>in</strong> Eq. (7.65) is to be replaced with −γ S . Its explicit<br />

dependence on the scale µ can be factored out via χ R (µ, D R ,τ R ,u R , q,ω)<br />

= µ −2 ˆχ R<br />

(<br />

τR ,u R , q/µ, ω/D R µ 2+a) , see Eq. (7.55), whence<br />

[<br />

−2 − γ S + γ D D R<br />

∂<br />

∂D R<br />

+ γ τ τ R<br />

∂<br />

∂τ R<br />

+ β u<br />

]<br />

∂<br />

ˆχ R (D R ,τ R ,u R )=0. (7.75)<br />

∂u R<br />

This l<strong>in</strong>ear partial differential equation is readily solved by means of the<br />

method of characteristics, as is Eq. (7.65) for the vertex functions. The idea


314 U. C. Täuber<br />

is to f<strong>in</strong>d a curve parametrisation µ(l) =µl <strong>in</strong> the space spanned by the<br />

parameters ˜D, ˜τ, andũ such that<br />

l d ˜D(l)<br />

dl<br />

= ˜D(l) γ D (l) , l d˜τ(l)<br />

dl<br />

= ˜τ(l) γ τ (l) , l dũ(l)<br />

dl<br />

= β u (l) , (7.76)<br />

with <strong>in</strong>itial values D R , τ R ,andu R , respectively at l =1.Thefirst-order ord<strong>in</strong>ary<br />

differential equations (7.76), with γ D (l) =γ D (ũ(l)) etc. def<strong>in</strong>e runn<strong>in</strong>g<br />

coupl<strong>in</strong>gs that describe how the parameters of the theory change under scale<br />

transformations µ → µl. The formal solutions for ˜D(l) and˜τ(l) read<br />

˜D(l) =D R exp<br />

[ ∫ l<br />

1<br />

γ D (l ′ ) dl′<br />

l ′ ]<br />

, ˜τ(l) =τ R exp<br />

[ ∫ l<br />

1<br />

γ τ (l ′ ) dl′<br />

l ′ ]<br />

. (7.77)<br />

For the function ˆχ(l) =ˆχ R ( ˜D(l), ˜τ(l), ũ(l)), we then obta<strong>in</strong> another ord<strong>in</strong>ary<br />

differential equation, namely<br />

which is solved by<br />

l dˆχ(l)<br />

dl<br />

ˆχ(l) =ˆχ(1) l 2 exp<br />

=[2+γ S (l)] ˆχ(l) , (7.78)<br />

[ ∫ l<br />

1<br />

γ S (l ′ ) dl′<br />

l ′ ]<br />

. (7.79)<br />

Collect<strong>in</strong>g everyth<strong>in</strong>g, we f<strong>in</strong>ally arrive at<br />

[<br />

χ R (µ, D R ,τ R ,u R , q,ω)=(µl) −2 exp<br />

× ˆχ R<br />

(<br />

−<br />

∫ l<br />

1<br />

γ S (l ′ ) dl′<br />

l ′ ]<br />

˜τ(l), ũ(l), |q|<br />

µl , ω<br />

˜D(l)(µl) 2+a )<br />

. (7.80)<br />

The solution (7.80) of the RG equation (7.75), along with the flow equations<br />

(7.76), (7.77) for the runn<strong>in</strong>g coupl<strong>in</strong>gs tell us how the dynamic susceptibility<br />

depends on the (momentum) scale µlat which we consider the theory.<br />

Similar relations can be obta<strong>in</strong>ed for arbitrary vertex functions by solv<strong>in</strong>g the<br />

associated RG equations (7.65) [13]. The po<strong>in</strong>t here is that the right-hand side<br />

of Eq. (7.80) may be evaluated outside the IR-s<strong>in</strong>gular regime, by fix<strong>in</strong>g one<br />

of its arguments at a f<strong>in</strong>ite value, say |q|/µ l = 1. The function ˆχ R is regular,<br />

and can be calculated by means of perturbation theory. A scale-<strong>in</strong>variant<br />

regime is characterised by the renormalised nonl<strong>in</strong>ear coupl<strong>in</strong>g u R becom<strong>in</strong>g<br />

<strong>in</strong>dependent of the scale µl,orũ(l) → u ∗ = const. For an RG fixed po<strong>in</strong>t to<br />

be <strong>in</strong>frared-stable, we thus require<br />

β u (u ∗ )=0, β ′ u(u ∗ ) > 0 , (7.81)


7 Field-Theory Approaches to Nonequilibrium Dynamics 315<br />

s<strong>in</strong>ce Eq. (7.76) then implies that ũ(l → 0) → u ∗ . Tak<strong>in</strong>g the limit l → 0<br />

thus provides the desired mapp<strong>in</strong>g of physical observables such as (7.80) onto<br />

the critical region. In the vic<strong>in</strong>ity of an IR-stable RG fixed po<strong>in</strong>t, Eq. (7.77)<br />

yields the power laws ˜D(l) ≈ D R l γ∗ D , where γ<br />

∗<br />

D = γ D (l → 0) = γ D (u ∗ ), etc.<br />

Consequently, Eq. (7.80) reduces to<br />

(<br />

χ R (τ R , q,ω) ≈ µ −2 l −2−γ∗ S ˆχR τ R l γ∗ τ ,u ∗ , |q|<br />

)<br />

µl , ω<br />

, (7.82)<br />

D R µ 2+a l 2+a+γ∗ D<br />

and upon match<strong>in</strong>g l = |q|/µ we recover the dynamic scal<strong>in</strong>g law (7.12) with<br />

the critical exponents<br />

η = −γ ∗ S , ν = −1/γ ∗ τ , z =2+a + γ ∗ D . (7.83)<br />

To one-loop order, we obta<strong>in</strong> from the RG beta function (7.74)<br />

u ∗ H =<br />

6 ɛ<br />

n +8 + O(ɛ2 ) . (7.84)<br />

Here we have <strong>in</strong>dicated that our perturbative expansion for small u has effectively<br />

turned <strong>in</strong>to a dimensional expansion <strong>in</strong> ɛ = d c − d. In dimensions<br />

d0. With<br />

Eqs. (7.71) and (7.73), the identifications (7.83) then give us explicit results<br />

for the static scal<strong>in</strong>g exponents, as mere functions of dimension d =4− ɛ and<br />

the number of order parameter components n,<br />

η = n +2<br />

2(n +8) 2 ɛ2 + O(ɛ 3 ) ,<br />

1<br />

ν<br />

=2−<br />

n +2<br />

n +8 ɛ + O(ɛ2 ) . (7.85)<br />

For model A with nonconserved order parameter, the two-loop result (7.72)<br />

yields the <strong>in</strong>dependent dynamic critical exponent<br />

a =0 : z =2+cη , c=6ln 4 − 1+O(ɛ) ; (7.86)<br />

3<br />

for model B with conserved order parameter, <strong>in</strong>stead γD ∗ = γ∗ S = −η, whence<br />

we arrive at the exact scal<strong>in</strong>g relation<br />

a =2 : z =4− η. (7.87)<br />

In dimensions d > d c = 4, the Gaussian fixed po<strong>in</strong>t u ∗ 0 = 0 is stable<br />

(β ′ u(0) = −ɛ >0). Therefore all anomalous dimensions disappear, i.e., γ ∗ S =<br />

0=γ ∗ D and γ∗ τ = −2, and we are left with the mean-field critical exponents<br />

η 0 =0,ν 0 =1/2, and z 0 =2+a. Precisely at the upper critical dimension<br />

d c = 4, the RG flow equation for the nonl<strong>in</strong>ear coupl<strong>in</strong>g becomes<br />

which is solved by<br />

l dũ(l)<br />

dl<br />

= n +8<br />

6<br />

(<br />

ũ(l) 2 + O ũ(l) 3) , (7.88)


316 U. C. Täuber<br />

ũ(l) =<br />

u R<br />

1 − n+8<br />

6<br />

u R ln l . (7.89)<br />

In four dimensions, ũ(l) → 0, but only logarithmically slowly, which causes<br />

logarithmic corrections to the mean-field critical power laws. For example,<br />

upon <strong>in</strong>sert<strong>in</strong>g Eq. (7.89) <strong>in</strong>to the flow equation (7.76), one f<strong>in</strong>ds ˜τ(l) ∼<br />

τ R l −2 (ln |l|) −(n+2)/(n+8) ; with ˜τ(l = ξ −1 )=O(1), iterative <strong>in</strong>version yields<br />

ξ(τ R ) ∼ τ −1/2<br />

R<br />

(ln τ R ) (n+2)/2(n+8) . (7.90)<br />

This concludes our derivation of asymptotic scal<strong>in</strong>g laws for the critical<br />

dynamics of the purely relaxational models A and B, and the explicit computation<br />

of the scal<strong>in</strong>g exponents <strong>in</strong> powers of ɛ = d c − d. In the follow<strong>in</strong>g<br />

sections, I will briefly sketch how the response functional formalism and the<br />

dynamic renormalisation group can be employed to study the critical dynamics<br />

of systems with reversible mode-coupl<strong>in</strong>g terms, the “age<strong>in</strong>g’ behaviour<br />

<strong>in</strong>duced by quench<strong>in</strong>g from random <strong>in</strong>itial conditions to the critical po<strong>in</strong>t, the<br />

effects of violat<strong>in</strong>g the detailed balance constra<strong>in</strong>ts on universal dynamic critical<br />

properties, and the generically scale-<strong>in</strong>variant features of nonequilibrium<br />

systems such as driven diffusive Is<strong>in</strong>g lattice gases.<br />

7.1.6 Critical Dynamics with Reversible Mode-Coupl<strong>in</strong>gs<br />

In the previous chapters, we have assumed purely relaxational dynamics for<br />

the order parameter, see Eq. (7.16). In general, however, there are also reversible<br />

contributions to the systematic force terms F α that enter its Langev<strong>in</strong><br />

equation [7,24]. Consider the Hamiltonian dynamics of microscopic variables,<br />

say, local sp<strong>in</strong> densities, at T =0:∂ t Sm(x, α t) = {H[S m ],Sm(x, α t)}. Here,<br />

the Poisson brackets {A, B} constitute the classical analog of the quantummechanical<br />

commutator<br />

i <br />

[A, B] (correspondence pr<strong>in</strong>ciple). Upon coarsegra<strong>in</strong><strong>in</strong>g,<br />

the microscopic variables Sm α become the mesoscopic hydrodynamic<br />

fields S α . S<strong>in</strong>ce the set of slow modes should provide a complete description<br />

of the critical dynamics, we may formally expand<br />

{<br />

}<br />

H[S],S α (x) =<br />

∫<br />

d d x ∑ ′ β<br />

δH[S]<br />

δS β (x ′ ) Qβα (x ′ , x) , (7.91)<br />

with the mutual Poisson brackets of the hydrodynamic variables<br />

{<br />

}<br />

Q αβ (x, x ′ )= S α (x),S β (x ′ ) = −Q βα (x ′ , x) . (7.92)<br />

By <strong>in</strong>spection of the associated Fokker–Planck equation, one may then<br />

establish an additional equilibrium condition <strong>in</strong> order for the time-dependent<br />

probability distribution to reach the canonical limit (7.6): P[S, t] →P eq [S] as<br />

t →∞provided the probability current is divergence-free <strong>in</strong> the space spanned<br />

by the stochastic fields S α (x):


7 Field-Theory Approaches to Nonequilibrium Dynamics 317<br />

∫<br />

d d x ∑ δ<br />

(<br />

δS α F α<br />

(x)<br />

rev[S] e −H[S]/kBT =0.<br />

α<br />

(7.93)<br />

It turns out that this equilibrium condition is often more crucial than the<br />

E<strong>in</strong>ste<strong>in</strong> relation (7.17). In order to satisfy Eq. (7.93) at T ≠ 0, we must<br />

supplement Eq. (7.91) by a f<strong>in</strong>ite-temperature correction, whereupon the reversible<br />

mode-coupl<strong>in</strong>g contributions to the systematic forces become<br />

∫<br />

Frev[S](x) α =− d d x ∑ ′ β<br />

[<br />

Q αβ (x, x ′ ) δH[S]<br />

δS β (x ′ ) − k BT δQαβ (x, x ′ ]<br />

)<br />

δS β (x ′ ,<br />

)<br />

(7.94)<br />

and the complete coupled set of stochastic differential equations reads<br />

∂S α (x,t)<br />

∂t<br />

= F α rev[S](x,t) − D α (i∇) aα δH[S]<br />

δS α (x,t) + ζα (x,t) , (7.95)<br />

where as before the D α denote the relaxation coefficients, and a α =0or2<br />

respectively for nonconserved and conserved modes.<br />

As an <strong>in</strong>structive example, let us consider the Heisenberg model for<br />

isotropic ferromagnets, H[{S j }]=− 1 ∑ N<br />

2 j,k=1 J jk S j · S k , where the sp<strong>in</strong> operators<br />

satisfy the usual commutation relations [Sj α,Sβ k ]=i ∑ γ ɛαβγ S γ j δ jk.<br />

The correspond<strong>in</strong>g Poisson brackets for the magnetisation density read<br />

Q αβ (x, x ′ )=−g ∑ γ<br />

ɛ αβγ S γ (x) δ(x − x ′ ) , (7.96)<br />

where the purely dynamical coupl<strong>in</strong>g g <strong>in</strong>corporates various factors that<br />

emerge upon coarse-gra<strong>in</strong><strong>in</strong>g and tak<strong>in</strong>g the cont<strong>in</strong>uum limit. The second<br />

contribution <strong>in</strong> Eq. (7.94) vanishes, s<strong>in</strong>ce it reduces to a contraction of the<br />

antisymmetric tensor ɛ αβγ with the Kronecker symbol δ βγ , whence we arrive<br />

at the Langev<strong>in</strong> equations govern<strong>in</strong>g the critical dynamics of the three order<br />

parameter components for isotropic ferromagnets [25]<br />

∂S(x,t)<br />

∂t<br />

= −g S(x,t) × δH[S]<br />

δS(x,t) + D δH[S]<br />

∇2 + ζ(x,t) , (7.97)<br />

δS(x,t)<br />

with 〈ζ(x,t)〉 = 0. S<strong>in</strong>ce [H[{S j }], ∑ k Sα k<br />

] = 0, the total magnetisation is<br />

conserved, whence the noise correlators should be taken as<br />

〈<br />

ζ α (x,t) ζ β (x ′ ,t ′ ) 〉 = −2Dk B T ∇ 2 δ(x − x ′ ) δ(t − t ′ ) δ αβ . (7.98)<br />

The vector product term <strong>in</strong> Eq. (7.97) describes the sp<strong>in</strong> precession <strong>in</strong> the local<br />

effective magnetic field δH[S]/δS, which <strong>in</strong>cludes a contribution <strong>in</strong>duced by<br />

the exchange <strong>in</strong>teraction.<br />

The Langev<strong>in</strong> equation (7.97) and (7.98) with the Hamiltonian (7.7) for<br />

n = 3 def<strong>in</strong>e the so-called model J [7]. In addition to the model B response


318 U. C. Täuber<br />

functional (7.35) and (7.36) with a = 2 (sett<strong>in</strong>g k B T = 1 aga<strong>in</strong>), the reversible<br />

force <strong>in</strong> Eq. (7.97) leads to an additional contribution to the action<br />

∫ ∫<br />

A mc [ ˜S,S]=−g d d x dt ∑ ɛ αβγ ˜Sα S β ( ∇ 2 S γ + h γ) , (7.99)<br />

α,β,γ<br />

which gives rise to an additional mode-coupl<strong>in</strong>g vertex, as depicted <strong>in</strong> Fig. 7.5(a).<br />

Power count<strong>in</strong>g yields the scal<strong>in</strong>g dimension [g] =µ 3−d/2 for the associated<br />

coupl<strong>in</strong>g strength, whence we expect a dynamical upper critical dimension<br />

d ′ c = 6. However, s<strong>in</strong>ce we are <strong>in</strong>vestigat<strong>in</strong>g a system <strong>in</strong> thermal equilibrium,<br />

we can treat its thermodynamics and static properties separately from its dynamics.<br />

Obviously therefore, the static critical exponents must still be given<br />

(to lowest nontrivial order and for d


7 Field-Theory Approaches to Nonequilibrium Dynamics 319<br />

where B d = Γ (4 − d/2)/2 d dπ d/2 , Eq. (7.101) implies the identity [9]<br />

Z g = Z S . (7.103)<br />

For the RG beta function associated with the effective coupl<strong>in</strong>g enter<strong>in</strong>g the<br />

loop corrections, we thus <strong>in</strong>fer<br />

β f = µ ∂<br />

∂µ ∣ f R = f R (d − 6+γ S − 2 γ D ) . (7.104)<br />

0<br />

Consequently, at any nontrivial IR-stable RG fixed po<strong>in</strong>t 0


320 U. C. Täuber<br />

7.1.7 Critical Relaxation, Initial Slip, and Age<strong>in</strong>g<br />

We beg<strong>in</strong> with a brief discussion of the coarsen<strong>in</strong>g dynamics of systems described<br />

by model A/B k<strong>in</strong>etics that are rapidly quenched from a disordered<br />

state at T ≫ T c to the critical po<strong>in</strong>t T ≈ T c [27, 28]. The situation may be<br />

modeled as a relaxation from Gaussian random <strong>in</strong>itial conditions, i.e., the<br />

probability distribution for the order parameter at t = 0 can be taken as<br />

(<br />

P[S, t =0]∝ e −H0[S] =exp − ∆ ∫<br />

d d x ∑ )<br />

[S α (x, 0) − a α (x)] 2 , (7.108)<br />

2<br />

α<br />

where the functions a α (x) specify the most likely <strong>in</strong>itial configurations. Power<br />

count<strong>in</strong>g for the parameter ∆ gives [∆] =µ 2 , whence it is a relevant perturbation<br />

that will flow to ∆ →∞under the RG. Asymptotically, therefore, the<br />

system will be governed by sharp Dirichlet boundary conditions. Whereas the<br />

response propagators rema<strong>in</strong>s a causal function of the time difference between<br />

applied perturbation and effect, G 0 (q,t−t ′ )=Θ(t−t ′ )e −Dqa (r+q 2 )(t−t ′) ,see<br />

Eq. (7.23), time translation <strong>in</strong>variance is broken by the <strong>in</strong>itial state <strong>in</strong> the<br />

Dirichlet correlator of the Gaussian model,<br />

C D (q; t, t ′ )= 1 (<br />

r + q 2 e −Dqa (r+q 2 ) |t−t ′| ) − e −Dqa (r+q 2 )(t+t ′ )<br />

. (7.109)<br />

Away from criticality, i.e., for r>0andq ≠ 0, temporal correlations decay<br />

exponentially fast, and the system quickly approaches the stationary equilibrium<br />

state. However, as T → T c , the equilibration time diverges accord<strong>in</strong>g to<br />

t c ∼|τ| −zν →∞, and the system never reaches thermal equilibrium. Twotime<br />

correlation functions will then depend on both times separately, <strong>in</strong> a<br />

specific manner to be addressed below, a phenomenon termed critical “age<strong>in</strong>g”<br />

(for more details, see Refs. [29, 30]).<br />

The field-theoretic treatment of the model A/B dynamical action (7.35),<br />

(7.36) with the <strong>in</strong>itial term (7.108) follows the theory of boundary critical<br />

phenomena [31]. However, it turns out that additional s<strong>in</strong>gularities on the<br />

temporal “surface” at t + t ′ = 0 appear only for model A, and can be <strong>in</strong>corporated<br />

<strong>in</strong>to a s<strong>in</strong>gle new renormalisation factor; to one-loop order, one<br />

f<strong>in</strong>ds [27, 28]<br />

a =0 : ˜Sα R (x, 0) = (Z 0 Z ˜S) 1/2 ˜Sα (x, 0) , Z 0 =1− n +2<br />

6<br />

u R<br />

ɛ<br />

. (7.110)<br />

This <strong>in</strong> turn leads to a s<strong>in</strong>gle <strong>in</strong>dependent critical exponent associated with<br />

the <strong>in</strong>itial time relaxation, the <strong>in</strong>itial slip exponent, which becomes for the<br />

purely relaxational models A and B with nonconserved and conserved order<br />

parameter:<br />

a =0 : θ = γ∗ 0<br />

2 z = n +2<br />

4(n +8) ɛ + O(ɛ2 ) , a =2 : θ =0. (7.111)


7 Field-Theory Approaches to Nonequilibrium Dynamics 321<br />

In order to obta<strong>in</strong> the short-time scal<strong>in</strong>g laws for the dynamic response and<br />

correlation functions <strong>in</strong> the age<strong>in</strong>g limit t ′ /t → 0, one requires additional<br />

<strong>in</strong>formation that can be garnered from the short-distance operator product<br />

expansion for the fields,<br />

t → 0: ˜S(x,t)=˜σ(t) ˜S0 (x) , S(x,t)=σ(t) ˜S 0 (x) . (7.112)<br />

Subsequent analysis then yields eventually [27, 28]<br />

χ(q; t, t ′ → 0) = |q| z−2+η ( t<br />

t ′ ) θ<br />

ˆχ 0 (q ξ,|q| z Dt) , (7.113)<br />

C(q; t, t ′ → 0) = |q| −2+η ( t<br />

t ′ ) θ−1<br />

Ĉ0 (q ξ,|q| z Dt) , (7.114)<br />

and for the time dependence of the mean order parameter<br />

( )<br />

〈S(t)〉 = S 0 t θ′ Ŝ S 0 t θ′ +β/zν<br />

, (7.115)<br />

a =0 : θ ′ = θ − z − 2+η<br />

z<br />

, a =2 : θ ′ = θ =0. (7.116)<br />

One may also compute the universal fluctuation–dissipation ratios <strong>in</strong> this<br />

nonequilibrium age<strong>in</strong>g regime [29, 30]. It emerges, though, that these depend<br />

on the quantity under <strong>in</strong>vestigation, which prohibits a unique def<strong>in</strong>ition of an<br />

effective nonequilibrium temperature for critical age<strong>in</strong>g. The method sketched<br />

above can be extended to models with reversible mode-coupl<strong>in</strong>gs [32]. For<br />

model J captur<strong>in</strong>g the critical dynamics of isotropic ferromagnets, one f<strong>in</strong>ds<br />

θ = z − 4+η<br />

z<br />

= − 6 − d − η<br />

d +2− η ; (7.117)<br />

<strong>in</strong> systems where a nonconserved order parameter is dynamically coupled to<br />

other conserved modes, the <strong>in</strong>itial slip exponent θ is actually not a universal<br />

number, but depends on the width of the <strong>in</strong>itial distribution [32].<br />

7.1.8 Nonequilibrium Relaxational Critical Dynamics<br />

Next we address the question [33], What happens if the detailed balance conditions<br />

(7.17) and (7.93) are violated? To start, we change the noise strength<br />

D → ˜D <strong>in</strong> the purely relaxational models A and B, which (<strong>in</strong> our units) violates<br />

the E<strong>in</strong>ste<strong>in</strong> relation (7.17). However, this modification can obviously be<br />

absorbed <strong>in</strong>to a rescaled effective temperature, k B T → k B T ′ = ˜D/D. Formally<br />

this is established by means of the dynamical action (7.34), which now reads<br />

∫ ∫<br />

A[ ˜S,S] = d d x dt ∑ ˜S<br />

[∂ α t S α + D (i∇) a ( r − ∇ 2) S α<br />

α<br />

− ˜D (i∇) a ˜Sα + D u 6 (i∇)a S ∑ ]<br />

α S β S β . (7.118)<br />

β


322 U. C. Täuber<br />

√<br />

Upon simple rescal<strong>in</strong>g ˜S α → ˜S ′α = ˜S α ˜D/D, S α → S ′α = S<br />

√D/ α ˜D, the<br />

response functional (7.118) recovers its equilibrium form, albeit with modified<br />

nonl<strong>in</strong>ear coupl<strong>in</strong>g u → ũ = u ˜D/D. However, the universal asymptotic<br />

properties of these models are governed by the Heisenberg fixed po<strong>in</strong>t (7.84),<br />

and the specific value of the (renormalised) coupl<strong>in</strong>g, which only serves as<br />

the <strong>in</strong>itial condition for the RG flow, does not matter. In fact, the relaxational<br />

dynamics of the k<strong>in</strong>etic Is<strong>in</strong>g model with Glauber dynamics (model<br />

A with n = 1) is known to be quite stable aga<strong>in</strong>st nonequilibrium perturbations<br />

[34, 35], even if these break the Is<strong>in</strong>g Z 2 symmetry [36]. For model J<br />

the above rescal<strong>in</strong>g modifies <strong>in</strong> a similar manner √ merely the mode-coupl<strong>in</strong>g<br />

strength <strong>in</strong> Eq. (7.99), namely g → ˜g = g ˜D/D [37]. Aga<strong>in</strong>, s<strong>in</strong>ce the dynamic<br />

critical behaviour is governed by the universal fixed po<strong>in</strong>t (7.107), thermal<br />

equilibrium becomes effectively restored at criticality. More generally, it<br />

has been established that isotropic detailed balance violations do not affect<br />

the universal properties <strong>in</strong> other models for critical dynamics that conta<strong>in</strong><br />

additional conserved variables either: the equilibrium RG fixed po<strong>in</strong>ts tend to<br />

be asymptotically stable [33].<br />

In systems with conserved order parameter, however, we may <strong>in</strong> addition<br />

<strong>in</strong>troduce spatially anisotropic violations of E<strong>in</strong>ste<strong>in</strong>’s relation; for example, <strong>in</strong><br />

model B one can allow for anisotropic relaxation −D ∇ 2 →−D ⊥ ∇ 2 ⊥ −D ‖ ∇ 2 ‖ ,<br />

with different rates <strong>in</strong> two spatial subsectors and concomitantly anisotropic<br />

noise correlations − ˜D ∇ 2 →−˜D ⊥ ∇ 2 ⊥ − ˜D ‖ ∇ 2 ‖. We have thus produced a<br />

truly nonequilibrium situation provided ˜D ⊥ /D ⊥ ≠ ˜D ‖ /D ‖ , which we may<br />

<strong>in</strong>terpret as hav<strong>in</strong>g effectively coupled the longitud<strong>in</strong>al and transverse spatial<br />

sectors to heat baths with different temperatures T ⊥


7 Field-Theory Approaches to Nonequilibrium Dynamics 323<br />

Quite remarkably, the Langev<strong>in</strong> equation (7.119) can be derived as an equilibrium<br />

diffusive relaxational k<strong>in</strong>etics<br />

∂S α (x,t)<br />

∂t<br />

= D ∇ 2 ⊥<br />

from an effective long-range Hamiltonian<br />

∫<br />

H eff [S] =<br />

+ ũ<br />

4!<br />

δH eff [S]<br />

δS α (x,t) + ζα (x,t) (7.121)<br />

d d q q 2 ⊥ (r + q2 ⊥ )+c q2 ∑<br />

‖<br />

∣ S α<br />

(2π) d 2 q 2 (q) ∣ 2<br />

⊥<br />

α<br />

∫<br />

d d x ∑ [S α (x)] 2 [S β (x)] 2 . (7.122)<br />

α,β<br />

Power count<strong>in</strong>g gives [ũ] =µ 4−d ‖−d : the spatial anisotropy suppresses longitud<strong>in</strong>al<br />

fluctuations and lower the upper critical dimension to d c =4−d ‖ .The<br />

anisotropic correlations encoded <strong>in</strong> Eq. (7.122) also reduce the lower critical<br />

dimension and affect the nature of the ordered phase [39, 41].<br />

The scal<strong>in</strong>g law for, e.g., the dynamic response function takes the form<br />

( √ )<br />

τ c<br />

χ(τ ⊥ , q ⊥ , q ‖ ,ω)=|q ⊥ | −2+η ˆχ<br />

|q ⊥ | , |q‖ |<br />

1/ν |q ⊥ | 1+∆ , ω<br />

D |q ⊥ | z , (7.123)<br />

where we have <strong>in</strong>troduced a new anisotropy exponent ∆. S<strong>in</strong>ce the nonl<strong>in</strong>ear<br />

coupl<strong>in</strong>g ũ only affects the transverse sector, we f<strong>in</strong>d to all orders <strong>in</strong> the<br />

perturbation expansion:<br />

Γ (1,1) (q ⊥ =0, q ‖ ,ω)=iω + Dcq 2 ‖ , (7.124)<br />

and consequently obta<strong>in</strong> the Z factor identity<br />

Z c = Z −1<br />

D<br />

= Z−1 S<br />

, (7.125)<br />

which at any IR-stable RG fixed po<strong>in</strong>t implies the exact scal<strong>in</strong>g relations<br />

z =4− η,<br />

∆=1− γ∗ c<br />

2 =1− η 2 = z 2 − 1 , (7.126)<br />

whereas the scal<strong>in</strong>g exponents for the longitud<strong>in</strong>al sector read<br />

z ‖ =<br />

z<br />

1+∆ =2, ν ‖ = ν (1 + ∆) = ν (4 − η) . (7.127)<br />

2<br />

As for the equilibrium model B, the only <strong>in</strong>dependent critical exponents to<br />

be determ<strong>in</strong>ed are η and ν. To one-loop order, only the comb<strong>in</strong>atorics of the<br />

Feynman diagrams (see Fig. 7.2) enters their explicit values, whence one f<strong>in</strong>ds<br />

for d


324 U. C. Täuber<br />

η =0+O(ɛ 2 ) ,<br />

1<br />

ν<br />

=2−<br />

n +2<br />

n +8 ɛ + O(ɛ2 ) , (7.128)<br />

albeit with different ɛ =4−d−d ‖ . To two-loop order, however, the anisotropy<br />

manifestly affects the evaluation of the loop contributions, and the value for<br />

η deviates from the expression <strong>in</strong> Eq. (7.85) [40].<br />

Interest<strong>in</strong>gly, an analogously constructed nonequilibrium two-temperature<br />

model J with reversible mode-coupl<strong>in</strong>g vertex cannot be cast <strong>in</strong>to a form that<br />

is equivalent to an equilibrium system, for ow<strong>in</strong>g to the emerg<strong>in</strong>g anisotropy,<br />

the condition (7.93) cannot be satisfied. A one-loop RG analysis yields a runaway<br />

flow, and no stable RG fixed po<strong>in</strong>t is found [38]. Similar behaviour ensues<br />

<strong>in</strong> other anisotropic nonequilibrium variants of critical dynamics models with<br />

conserved order parameter; the precise <strong>in</strong>terpretation of the apparent <strong>in</strong>stability<br />

is as yet unclear [33].<br />

7.1.9 Driven Diffusive Systems<br />

F<strong>in</strong>ally, we wish to consider Langev<strong>in</strong> representations of genu<strong>in</strong>ely nonequilibrium<br />

systems, namely driven diffusive lattice gases (for a comprehensive<br />

overview, see Ref. [42]). First we address the coarse-gra<strong>in</strong>ed cont<strong>in</strong>uum version<br />

of the asymmetric exclusion process, i.e., hard-core repulsive particles that hop<br />

preferentially <strong>in</strong> one direction. We describe this system <strong>in</strong> terms of a conserved<br />

particle density, whose fluctuations we denote with S(x,t), such that 〈S〉 =0,<br />

obey<strong>in</strong>g a cont<strong>in</strong>uity equation ∂ t S(x,t)+∇ · J(x,t) = 0. We assume the system<br />

to be driven along the “‖’ direction; <strong>in</strong> the transverse sector (of dimension<br />

d ⊥ = d − 1) we thus just have a noisy diffusion current J ⊥ = −D ∇ ⊥ S + η,<br />

whereas there is a nonl<strong>in</strong>ear term, stemm<strong>in</strong>g from the hard-core <strong>in</strong>teractions,<br />

<strong>in</strong> the current along the direction of the external drive, with J 0 ‖ =const.: J ‖ =<br />

J 0 ‖ −Dc∇ ‖ S − 1 2 Dg S2 +η ‖ . For the stochastic currents, we assume Gaussian<br />

white noise 〈η i 〉 =0=〈η ‖ 〉 and 〈η i (x,t) η j (x ′ ,t ′ )〉 =2Dδ(x − x ′ ) δ(t − t ′ ) δ ij ,<br />

〈<br />

η‖ (x,t) η ‖ (x ′ ,t ′ ) 〉 =2D ˜cδ(x − x ′ ) δ(t − t ′ ). Notice that s<strong>in</strong>ce we are not <strong>in</strong><br />

thermal equilibrium, E<strong>in</strong>ste<strong>in</strong>’s relation need not be fulfilled. We can however<br />

always rescale the field to satisfy it <strong>in</strong> the transverse sector; the ratio w =˜c/c<br />

then measures the deviation from equilibrium. These considerations yield the<br />

generic Langev<strong>in</strong> equation for the density fluctuations <strong>in</strong> driven diffusive systems<br />

(DDS) [43, 44]<br />

∂S(x,t)<br />

∂t<br />

( )<br />

= D ∇ 2 ⊥ + c ∇ 2 ‖ S(x,t)+ Dg<br />

2 ∇ ‖S(x,t) 2 + ζ(x,t) , (7.129)<br />

with conserved noise ζ = −∇ ⊥ · η −∇ ‖ η ‖ , where 〈ζ〉 =0and<br />

( )<br />

〈ζ(x,t) ζ(x ′ ,t ′ )〉 = −2D ∇ 2 ⊥ +˜c ∇ 2 ‖ δ(x − x ′ ) δ(t − t ′ ) . (7.130)<br />

Notice that the drive term ∝ g breaks both the system’s spatial reflection<br />

symmetry and the Is<strong>in</strong>g Z 2 symmetry S →−S.


7 Field-Theory Approaches to Nonequilibrium Dynamics 325<br />

The correspond<strong>in</strong>g Janssen–De Dom<strong>in</strong>icis response functional (7.34) reads<br />

∫ ∫ [<br />

A[ ˜S,S] = d d x dt ˜S ∂S<br />

( )<br />

∂t − D ∇ 2 ⊥ + c ∇ 2 ‖ S<br />

]<br />

( )<br />

+ D ∇ 2 ⊥ +˜c ∇ 2 Dg<br />

‖ ˜S −<br />

2 ∇ ‖ S 2 . (7.131)<br />

It describes a “massless” theory, hence we expect the system to be generically<br />

scale-<strong>in</strong>variant, without the need to tune it to a special po<strong>in</strong>t <strong>in</strong> parameter<br />

space. The nonl<strong>in</strong>ear drive term will <strong>in</strong>duce anomalous scal<strong>in</strong>g <strong>in</strong> the drive<br />

direction, different from ord<strong>in</strong>ary diffusive behaviour. In the transverse sector,<br />

however, we have to all orders <strong>in</strong> the perturbation expansion simply<br />

Γ (1,1) (q ⊥ ,q ‖ =0,ω)=iω +D q 2 ⊥ , Γ (2,0) (q ⊥ ,q ‖ =0,ω)=−2D q 2 ⊥ , (7.132)<br />

s<strong>in</strong>ce the nonl<strong>in</strong>ear three-po<strong>in</strong>t vertex, which is of the form depicted <strong>in</strong><br />

Fig. 7.5(a), is proportional to iq ‖ . Consequently,<br />

which immediately implies<br />

Z ˜S<br />

= Z S = Z D =1, (7.133)<br />

η =0, z =2. (7.134)<br />

Moreover, the nonl<strong>in</strong>ear coupl<strong>in</strong>g g itself does not renormalise either as a<br />

consequence of Galilean <strong>in</strong>variance. Namely, the Langev<strong>in</strong> equation (7.129)<br />

and the action (7.131) are left <strong>in</strong>variant under Galilean transformations<br />

S ′ (x ′ ⊥,x ′ ‖ ,t′ )=S(x ⊥ ,x ‖ − Dgv t, t) − v ; (7.135)<br />

thus, the boost velocity v must scale as the field S under renormalisation, and<br />

s<strong>in</strong>ce the product Dgv must be <strong>in</strong>variant under the RG, this leaves us with<br />

Z g = Z −1<br />

D<br />

Z−1 S<br />

=1. (7.136)<br />

The effective nonl<strong>in</strong>ear coupl<strong>in</strong>g govern<strong>in</strong>g the perturbation expansion <strong>in</strong><br />

terms of loop diagrams turns out to be g 2 /c 3/2 ; if we def<strong>in</strong>e its renormalised<br />

counterpart as<br />

v R = Zc 3/2 vC d µ d−2 , (7.137)<br />

with the convenient choice C d = Γ (2 − d/2)/2 d−1 π d/2 , we see that the associated<br />

RG beta function becomes<br />

β v = v R<br />

(d − 2 − 3 )<br />

2 γ c . (7.138)<br />

At any nontrivial RG fixed po<strong>in</strong>t 0 < v ∗ < ∞, therefore γ ∗ c = 2 3<br />

(d − 2).<br />

We thus <strong>in</strong>fer that below the upper critical dimension d c = 2 for DDS, the<br />

longitud<strong>in</strong>al scal<strong>in</strong>g exponents are fixed by the system’s symmetry [43, 44],


326 U. C. Täuber<br />

∆ = − γ∗ c<br />

2 = 2 − d<br />

3<br />

, z ‖ = 2<br />

1+∆ = 6<br />

5 − d . (7.139)<br />

An explicit one-loop calculation for the two-po<strong>in</strong>t vertex functions, see<br />

Fig. 7.5(b) and (c), yields<br />

γ c = − v R<br />

16 (3 + w R) , γ˜c = − v R<br />

(<br />

3w<br />

−1<br />

R<br />

32<br />

R)<br />

, (7.140)<br />

β w = w R (γ˜c − γ c )=− v R<br />

32 (w R − 1) (w R − 3) . (7.141)<br />

This establishes that <strong>in</strong> fact the fixed po<strong>in</strong>t w ∗ = 1 is IR-stable (provided 0 <<br />

v ∗ < ∞), which means that asymptotically the E<strong>in</strong>ste<strong>in</strong> relation is satisfied<br />

<strong>in</strong> the longitud<strong>in</strong>al sector as well [43].<br />

In this context, it is <strong>in</strong>structive to make an <strong>in</strong>trigu<strong>in</strong>g connection with the<br />

noisy Burgers equation [45], describ<strong>in</strong>g simplified fluid dynamics <strong>in</strong> terms of<br />

a velocity field u(x,t):<br />

∂u(x,t)<br />

+ Dg<br />

∂t 2 ∇ [ u(x,t) 2] = D ∇ 2 u(x,t)+ζ(x,t) , (7.142)<br />

〈ζ i 〉 =0, 〈ζ i (x,t) ζ j (x ′ ,t ′ )〉 = −2D ∇ i ∇ j δ(x − x ′ ) δ(t − t ′ ) . (7.143)<br />

For Dg = 1, the nonl<strong>in</strong>earity is just the usual fluid advection term. In one<br />

dimension, the Burgers equation (7.142) becomes identical with the DDS<br />

Langev<strong>in</strong> equation (7.129), so we immediately <strong>in</strong>fer its anomalous dynamic<br />

critical exponent z ‖ =3/2. At least <strong>in</strong> one dimension therefore, it should represent<br />

an equilibrium system which asymptotically approaches the canonical<br />

distribution (7.6), where the Hamiltonian is simply the fluid’s k<strong>in</strong>etic energy<br />

(andwehavesetk B T = 1). So let us check the equilibrium condition (7.93)<br />

with P eq [u] ∝ exp [ ∫<br />

− 1 2 u(x) 2 d d x ] :<br />

∫<br />

d d δ<br />

x<br />

δu(x,t) · [∇u(x,t) 2] e − 1 ∫<br />

2 u(x ′ ,t) 2 d d x ′<br />

=<br />

∫ [2∇<br />

· u(x,t) − u(x,t) · ∇u(x,t)<br />

2 ] d d x e − 1 2<br />

∫<br />

u(x ′ ,t) 2 d d x ′ .<br />

With appropriate boundary conditions, the first term here vanishes, but the<br />

second one does so only <strong>in</strong> d = 1: − ∫ u (du 2 /dx)dx = ∫ ∫ u 2 (du/dx)dx =<br />

1<br />

3 (du 3 /dx)dx = 0. Driven diffusive systems <strong>in</strong> one dimension are therefore<br />

subject to a “hidden” fluctuation–dissipation theorem.<br />

To conclude this part on Langev<strong>in</strong> dynamics, let us briefly consider the<br />

driven model B or critical DDS [42], which corresponds to a driven Is<strong>in</strong>g lattice<br />

gas near its critical po<strong>in</strong>t. Here, a conserved scalar field S undergoes a secondorder<br />

phase transition, but similar to the randomly driven case, aga<strong>in</strong> only the<br />

transverse sector is critical. Upon add<strong>in</strong>g the DDS drive term from Eq. (7.129)<br />

to the Langev<strong>in</strong> equation (7.119), we obta<strong>in</strong>


∂S(x,t)<br />

∂t<br />

7 Field-Theory Approaches to Nonequilibrium Dynamics 327<br />

[<br />

= D ∇ 2 ( ) ]<br />

⊥ r − ∇<br />

2<br />

⊥ + c ∇<br />

2<br />

‖ S(x,t)+ D ũ<br />

6 ∇2 ⊥ S(x,t) 3<br />

+ Dg<br />

2 ∇ ‖ S(x,t) 2 + ζ(x,t) , (7.144)<br />

with the (scalar) noise specified <strong>in</strong> Eq. (7.120). The response functional thus<br />

becomes<br />

∫ ∫ [<br />

A[ ˜S,S] = d d x dt ˜S ∂S<br />

[<br />

∂t − D ∇ 2 ( ) ]<br />

⊥ r − ∇<br />

2<br />

⊥ + c ∇<br />

2<br />

‖ S<br />

(<br />

+D ∇ 2 ˜S ⊥ − ũ<br />

6 ∇2 ⊥ S 3 − g ) ]<br />

2 ∇ ‖ S 2 . (7.145)<br />

Power count<strong>in</strong>g gives [g 2 ] = µ 5−d , so the upper critical dimension here is<br />

d c = 5, and [ũ] =µ 3−d . The nonl<strong>in</strong>earity ∝ ũ is thus irrelevant and can<br />

be omitted if we wish to determ<strong>in</strong>e the asymptotic universal scal<strong>in</strong>g laws;<br />

but recall that it is responsible for the phase transition <strong>in</strong> the system. The<br />

rema<strong>in</strong><strong>in</strong>g vertex is then proportional to iq ‖ , whence Eqs. (7.132) and (7.133)<br />

hold for critical DDS as well, and the transverse critical exponents are just<br />

those of the Gaussian model B,<br />

η =0, ν =1/2 , z =4. (7.146)<br />

In addition, Galilean <strong>in</strong>variance with respect to Eq. (7.135) and therefore<br />

Eq. (7.136) hold as before. With the renormalised nonl<strong>in</strong>ear drive strength<br />

def<strong>in</strong>ed similarly to Eq. (7.137), but a different geometric constant and the<br />

scale factor µ d−5 , the associated RG beta function reads<br />

β v = v R<br />

(d − 5 − 3 )<br />

2 γ c , (7.147)<br />

which aga<strong>in</strong> allows us to determ<strong>in</strong>e the longitud<strong>in</strong>al scal<strong>in</strong>g exponents to all<br />

orders <strong>in</strong> perturbation theory, for d


328 U. C. Täuber<br />

7.2 Reaction–Diffusion Systems<br />

We now turn our attention to stochastic <strong>in</strong>teract<strong>in</strong>g particle systems, whose<br />

microscopic dynamics is def<strong>in</strong>ed through a (classical) master equation. Below,<br />

we shall see how the latter can be mapped onto a stochastic quasi-Hamiltonian<br />

<strong>in</strong> a second-quantised bosonic operator representation [10–12]. Tak<strong>in</strong>g the<br />

cont<strong>in</strong>uum limit on the basis of coherent-state path <strong>in</strong>tegrals then yields a field<br />

theory action that may be analysed by the very same RG methods as described<br />

before <strong>in</strong> Subsects. 7.1.3–7.1.5 (for more details, see the recent overview [12]).<br />

7.2.1 Chemical Reactions and Population Dynamics<br />

Our goal is to study systems of “particles” A,B,... that propagate through<br />

hopp<strong>in</strong>g to nearest neighbors on a d-dimensional lattice, or via diffusion <strong>in</strong><br />

the cont<strong>in</strong>uum. Upon encounter, or spontaneously, with given stochastic rates,<br />

these particles may undergo species changes, annihilate, or produce offspr<strong>in</strong>g.<br />

At large densities, the characteristic time scales of the k<strong>in</strong>etics will be governed<br />

by the reaction rates, and the system is said to be reaction-limited. In contrast,<br />

at low densities, any reactions that require at least two particles to be <strong>in</strong><br />

proximity will be diffusion-limited: the basic time scale will be set by the<br />

hopp<strong>in</strong>g rate or diffusion coefficient.<br />

As a first approximation to the dynamics of such “chemical” reactions,<br />

let us assume homogeneous mix<strong>in</strong>g of each species. We may then hope to be<br />

able to capture the k<strong>in</strong>etics <strong>in</strong> terms of rate equations for each particle concentration<br />

or mean density. Note that such a description neglects any spatial<br />

fluctuations and correlations <strong>in</strong> the system, and is therefore <strong>in</strong> character a<br />

mean-field approximation. As a first illustration consider the annihilation of<br />

k−l >0 particles of species A <strong>in</strong> the irreversible kth-order reaction kA→ lA,<br />

with rate λ. The correspond<strong>in</strong>g rate equation employs a factorisation of the<br />

probability of encounter<strong>in</strong>g k particles at the same po<strong>in</strong>t to simply the kth<br />

power of the concentration a(t),<br />

ȧ(t) =−(k − l) λa(t) k . (7.149)<br />

This ord<strong>in</strong>ary differential equation is readily solved, with the result<br />

k =1 : a(t) =a(0) e −λt , (7.150)<br />

k ≥ 2: a(t) = [ a(0) 1−k +(k − l)(k − 1) λt ] −1/(k−1)<br />

. (7.151)<br />

For simple “radioactive” decay (k = 1), we of course obta<strong>in</strong> an exponential<br />

time dependence, as appropriate for statistically <strong>in</strong>dependent events. For<br />

pair (k = 2) and higher-order (k ≥ 3) processes, however, we f<strong>in</strong>d algebraic<br />

long-time behaviour, a(t) → (λt) −1/(k−1) , with an amplitude that becomes<br />

<strong>in</strong>dependent of the <strong>in</strong>itial density a(0). The absence of a characteristic time<br />

scale h<strong>in</strong>ts at cooperative effects, and we have to ask if and under which circumstances<br />

correlations might qualitatively affect the asymptotic long-time


7 Field-Theory Approaches to Nonequilibrium Dynamics 329<br />

power laws. For accord<strong>in</strong>g to Smoluchowski theory [12], we would expect the<br />

annihilation reactions to produce depletion zones <strong>in</strong> sufficiently low dimensions<br />

d ≤ d c , which would <strong>in</strong> turn <strong>in</strong>duce a considerable slow<strong>in</strong>g down of the<br />

density decay, see Subsect. 7.2.3. For two-species pair annihilation A + B →∅<br />

(without mix<strong>in</strong>g), another complication emerges, namely particle species segregation<br />

<strong>in</strong> dimensions for d ≤ d s ; the regions dom<strong>in</strong>ated by either species<br />

become largely <strong>in</strong>ert, and the annihilation reactions are conf<strong>in</strong>ed to rather<br />

sharp fronts [12].<br />

Competition between particle decay and production processes, e.g., <strong>in</strong> the<br />

reactions A →∅(with rate κ), A ⇋ A + A (with forward and back rates σ<br />

and λ, respectively), leads to even richer scenarios, as can already be <strong>in</strong>ferred<br />

from the associated rate equation<br />

ȧ(t) =(σ − κ) a(t) − λa(t) 2 . (7.152)<br />

For σκ, we encounter an<br />

active state with a(t) → a ∞ =(σ − κ)/λ exponentially, with rate ∼ σ − κ. We<br />

have thus identified a nonequilibrium cont<strong>in</strong>uous phase transition at σ c = κ.<br />

Indeed, as <strong>in</strong> equilibrium critical phenomena, the critical po<strong>in</strong>t is governed by<br />

characteristic power laws; for example, the asymptotic particle density a ∞ ∼<br />

(σ−σ c ) β , and the critical density decay a(t) ∼ (λt) −α with β 0 =1=α 0 <strong>in</strong> the<br />

mean-field approximation. The follow<strong>in</strong>g natural questions then arise: What<br />

are the critical exponents once statistical fluctuations are properly <strong>in</strong>cluded<br />

<strong>in</strong> the analysis? Can we, as <strong>in</strong> equilibrium systems, identify and characterise<br />

certa<strong>in</strong> universality classes, and which microscopic or overall, global features<br />

determ<strong>in</strong>e them and their critical dimension?<br />

Already the previous set of reactions may also be viewed as a (crude) model<br />

for the population dynamics of a s<strong>in</strong>gle species. In the same language, we may<br />

also formulate a stochastic version of the classic Lotka–Volterra predator–prey<br />

competition model [46]: if by themselves, the “predators” A die out accord<strong>in</strong>g<br />

to A →∅, with rate κ, whereas the prey reproduce B → B + B with rate σ,<br />

and thus proliferate with a Malthusian population explosion. The predators<br />

are kept alive and the prey under control through predation, here modeled as<br />

the reaction A + B → A + A: with rate λ, a prey is “eaten” by a predator,<br />

who simultaneously produces an offspr<strong>in</strong>g. The coupled k<strong>in</strong>etic rate equations<br />

for this system read<br />

ȧ(t) =λa(t) b(t) − κa(t) , ḃ(t) =σb(t) − λa(t) b(t) . (7.153)<br />

It is straightforward to show that the quantity K(t) =λ[a(t)+b(t)]−σ ln a(t)−<br />

κ ln b(t) is a constant of motion for this coupled system of differential equations,<br />

i.e., ˙K(t) = 0. As a consequence, the system is governed by regular<br />

population oscillations, whose frequency and amplitude are fully determ<strong>in</strong>ed


330 U. C. Täuber<br />

by the <strong>in</strong>itial conditions. Clearly, this is not a very realistic feature (albeit<br />

mathematically appeal<strong>in</strong>g), and moreover Eqs. (7.153) are known to be quite<br />

unstable with respect to model modifications [46]. Indeed, if one <strong>in</strong>cludes<br />

spatial degrees of freedom and takes account of the full stochasticity of the<br />

processes <strong>in</strong>volved, the system’s behaviour turns out to be much richer [47]:<br />

In the species coexistence phase, one encounters for sufficiently large values of<br />

the predation rate an <strong>in</strong>cessant sequence of “pursuit and evasion” waves that<br />

form quite complex dynamical patterns. In f<strong>in</strong>ite systems, these <strong>in</strong>duce erratic<br />

population oscillations whose features are however <strong>in</strong>dependent of the <strong>in</strong>itial<br />

configuration, but whose amplitude vanishes <strong>in</strong> the thermodynamic limit.<br />

Moreover, if locally the prey “carry<strong>in</strong>g capacity” is limited (correspond<strong>in</strong>g to<br />

restrict<strong>in</strong>g the maximum site occupation number per site on a lattice), there<br />

appears an ext<strong>in</strong>ction threshold for the predator population that separates the<br />

absorb<strong>in</strong>g state of a system filled with prey from the active coexistence regime<br />

through a cont<strong>in</strong>uous phase transition [47].<br />

These examples all call for a systematic approach to <strong>in</strong>clude stochastic<br />

fluctuations <strong>in</strong> the mathematical description of <strong>in</strong>teract<strong>in</strong>g reaction–diffusion<br />

systems that would be conducive to the application of field-theoretic tools, and<br />

thus allow us to br<strong>in</strong>g the powerful mach<strong>in</strong>ery of the dynamic renormalisation<br />

group to bear on these problems. In the follow<strong>in</strong>g, we shall describe such a<br />

general method [48–50] which allows a representation of the classical master<br />

equation <strong>in</strong> terms of a coherent-state path <strong>in</strong>tegral and its subsequent analysis<br />

by means of the RG (for overviews, see Refs. [10–12]).<br />

7.2.2 Field Theory Representation of Master Equations<br />

The above <strong>in</strong>teract<strong>in</strong>g particle systems, when def<strong>in</strong>ed on a d-dimensional lattice<br />

with sites i, are fully characterised by the set of occupation <strong>in</strong>teger numbers<br />

n i =0, 1, 2,... for each particle species. The master equation then describes<br />

the temporal evolution of the configurational probability distribution<br />

P ({n i }; t) through a balance of ga<strong>in</strong> and loss terms. For example, for the b<strong>in</strong>ary<br />

annihilation and coagulation reactions A + A →∅with rate λ and A + A → A<br />

with rate λ ′ , the master equation on a specific site i reads<br />

∂P(n i ; t)<br />

∂t<br />

= λ (n i +2)(n i +1)P (...,n i +2,...; t)<br />

+λ ′ (n i +1)n i P (...,n i +1,...; t)<br />

−(λ + λ ′ ) n i (n i − 1) P (...,n i ,...; t) , (7.154)<br />

with <strong>in</strong>itially P ({n i }, 0) = ∏ i P (n i), e.g., a Poisson distribution P (n i ) =<br />

¯n ni<br />

0 e−¯n0 /n i !. S<strong>in</strong>ce the reactions all change the site occupation numbers by<br />

<strong>in</strong>teger values, a second-quantised Fock space representation is particularly<br />

useful [48–50]. To this end, we <strong>in</strong>troduce the bosonic operator algebra<br />

] [ ] [ ]<br />

[a i ,a j =0= a † i ,a† j , a i ,a † j = δ ij . (7.155)


7 Field-Theory Approaches to Nonequilibrium Dynamics 331<br />

From these commutation relations one establishes <strong>in</strong> the standard manner<br />

that a i and a † i constitute lower<strong>in</strong>g and rais<strong>in</strong>g ladder operators, from which<br />

we may construct the particle number eigenstates |n i 〉,<br />

a i |n i 〉 = n i |n i − 1〉 , a † i |n i〉 = |n i +1〉 , a † i a i |n i 〉 = n i |n i 〉 . (7.156)<br />

(Notice that we have chosen a different normalisation than <strong>in</strong> ord<strong>in</strong>ary quantum<br />

mechanics.) A state with n i particles on sites i is then obta<strong>in</strong>ed from the<br />

empty vaccum state |0〉, def<strong>in</strong>ed through a i |0〉 = 0, as the product state<br />

|{n i }〉 = ∏ ( )<br />

a † ni<br />

i |0〉 . (7.157)<br />

i<br />

To make contact with the time-dependent configuration probability, we<br />

<strong>in</strong>troduce the formal state vector<br />

|Φ(t)〉 = ∑ {n i}<br />

P ({n i }; t) |{n i }〉 , (7.158)<br />

whereupon the l<strong>in</strong>ear time evolution accord<strong>in</strong>g to the master equation is translated<br />

<strong>in</strong>to an “imag<strong>in</strong>ary-time” Schröd<strong>in</strong>ger equation<br />

∂|Φ(t)〉<br />

∂t<br />

= −H |Φ(t)〉 , |Φ(t)〉 =e −Ht |Φ(0)〉 . (7.159)<br />

The stochastic quasi-Hamiltonian (rather, the time evolution or Liouville operator)<br />

for the on-site reaction processes is a sum of local terms, H reac =<br />

∑<br />

i H i(a † i ,a i); e.g., for the b<strong>in</strong>ary annihilation and coagulation reactions,<br />

(<br />

H i (a † i ,a i)=−λ 1 − a † 2 )<br />

i a 2 i − λ ′( )<br />

1 − a † i a † i a2 i . (7.160)<br />

The two contributions for each process may be physically <strong>in</strong>terpreted as follows:<br />

The first term corresponds to the actual process under consideration,<br />

and describes how many particles are annihilated and (re-)created <strong>in</strong> each<br />

reaction. The second term gives the “order” of each reaction, i.e., the number<br />

operator a † i a i appears to the kth power, but <strong>in</strong> normal-ordered form as<br />

a † k<br />

i a<br />

k<br />

i ,forakth-order process. Note that the reaction Hamiltonians such as<br />

(7.160) are non-Hermitean, reflect<strong>in</strong>g the particle creations and destructions.<br />

In a similar manner, hopp<strong>in</strong>g between neighbour<strong>in</strong>g sites 〈ij〉 is represented<br />

<strong>in</strong> this formalism through<br />

(<br />

)<br />

a † i − a† j<br />

)(a i − a j . (7.161)<br />

H diff = D ∑ 〈ij〉<br />

Our goal is of course to compute averages with respect to the configurational<br />

probability distribution P ({n i }; t); this is achieved by means of the<br />

projection state 〈P| = 〈0| ∏ i eai , which satisfies 〈P|0〉 =1and〈P|a † i = 〈P|,


332 U. C. Täuber<br />

s<strong>in</strong>ce [e ai ,a † j ]=eai δ ij . For the desired statistical averages of observables that<br />

must be expressible <strong>in</strong> terms of the occupation numbers {n i }, we then obta<strong>in</strong><br />

〈F (t)〉 = ∑ {n i}<br />

F ({n i }) P ({n i }; t) =〈P| F ({a † i a i}) |Φ(t)〉 . (7.162)<br />

Let first us explore the consequences of probability conservation, i.e., 1 =<br />

〈P|Φ(t)〉 = 〈P|e −Ht |Φ(0)〉. This requires 〈P|H = 0; upon commut<strong>in</strong>g e ∑ i ai<br />

with H, effectively the creation operators become shifted a † i → 1+a† i , whence<br />

this condition is fulfilled provided H i (a † i → 1,a i) = 0, which is <strong>in</strong>deed satisfied<br />

by our explicit expressions (7.160) and (7.161). By this prescription, we may<br />

also <strong>in</strong> averages replace a † i a i → a i , i.e., the particle density becomes a(t) =<br />

〈a i 〉, and the two-po<strong>in</strong>t operator a † i a i a † j a j → a i δ ij + a i a j .<br />

In the bosonic operator representation above, we have assumed that there<br />

exist no restrictions on the particle occupation numbers n i on each site. If,<br />

however, there is a maximum n i ≤ 2s+1, one may <strong>in</strong>stead employ a representation<br />

<strong>in</strong> terms of sp<strong>in</strong> s operators. For example, particle exclusion systems<br />

with n i = 0 or 1 can thus be mapped onto non-Hermitean sp<strong>in</strong> 1/2 “quantum”<br />

systems. Specifically <strong>in</strong> one dimension, such representations <strong>in</strong> terms of <strong>in</strong>tegrable<br />

sp<strong>in</strong> cha<strong>in</strong>s have proved a fruitful tool; for overviews, see Refs. [51–54].<br />

An alternative approach uses the bosonic theory, but encodes the site occupation<br />

restrictions through appropriate exponentials <strong>in</strong> the number operators<br />

e −a† i ai [55].<br />

We may now follow an established route <strong>in</strong> quantum many-particle theory<br />

[56] and proceed towards a field theory representation through construct<strong>in</strong>g<br />

the path <strong>in</strong>tegral equivalent to the “Schröd<strong>in</strong>ger” dynamics (7.159) based<br />

on coherent states, which are right eigenstates of the annihilation operator,<br />

a i |φ i 〉 = φ i |φ i 〉, with complex eigenvalues φ i . Explicitly, one f<strong>in</strong>ds<br />

(<br />

|φ i 〉 =exp − 1 )<br />

2 |φ i| 2 + φ i a † i |0〉 , (7.163)<br />

satisfy<strong>in</strong>g the overlap and (over-)completeness relations<br />

〈φ j |φ i 〉 =exp<br />

(− 1 2 |φ i| 2 − 1 ) ∫ ∏<br />

2 |φ j| 2 + φ ∗ d 2 φ i<br />

j φ i ,<br />

π |{φ i}〉 〈{φ i }| =1.<br />

i<br />

(7.164)<br />

Upon splitt<strong>in</strong>g the temporal evolution (7.159) <strong>in</strong>to <strong>in</strong>f<strong>in</strong>itesimal steps, and<br />

<strong>in</strong>sert<strong>in</strong>g Eq. (7.164) at each time step, standard procedures (elaborated <strong>in</strong><br />

detail <strong>in</strong> Ref. [12]) yield eventually<br />

∫ ∏<br />

〈F (t)〉 ∝<br />

i<br />

D[φ i ] D[φ ∗ i ] F ({φ i })e −A[φ∗ i ,φi] , (7.165)<br />

with the action


A[φ ∗ i ,φ i ]= ∑ i<br />

7 Field-Theory Approaches to Nonequilibrium Dynamics 333<br />

( ∫ tf<br />

[<br />

] )<br />

−φ i (t f )+ dt φ ∗ ∂φ i<br />

i<br />

0 ∂t + H i(φ ∗ i ,φ i ) −¯n 0 φ ∗ i (0) , (7.166)<br />

where the first term orig<strong>in</strong>ates from the projection state, and the last one from<br />

the <strong>in</strong>itial Poisson distribution. Notice that <strong>in</strong> the Hamiltonian, the creation<br />

and annihilation operators a † i and a i are simply replaced with the complex<br />

numbers φ ∗ i and φ i, respectively.<br />

Tak<strong>in</strong>g the cont<strong>in</strong>uum limit, φ i (t) → ψ(x,t), φ ∗ i (t) → ˆψ(x,t), the “bulk”<br />

part of the action becomes<br />

∫ ∫ [ ( )<br />

]<br />

A[ ˆψ, ∂<br />

ψ] = d d x dt ˆψ<br />

∂t − D ∇2 ψ + H reac ( ˆψ, ψ) , (7.167)<br />

where the hopp<strong>in</strong>g term (7.161) has naturally turned <strong>in</strong>to a diffusion propagator.<br />

We have thus arrived at a microscopic stochastic field theory for reaction–<br />

diffusion processes, with no assumptions whatsoever on the form of the (<strong>in</strong>ternal)<br />

noise. This is a crucial <strong>in</strong>gredient for nonequilibrium dynamics, and<br />

we may now use Eq. (7.167) as a basis for systematic coarse-gra<strong>in</strong><strong>in</strong>g and the<br />

renormalisation group analysis. Return<strong>in</strong>g to our example of pair annihilation<br />

and coagulation, the reaction part of the action (7.167) reads<br />

H reac ( ˆψ,<br />

(<br />

ψ) =−λ 1 − ˆψ 2) ( ˆψ)<br />

ψ 2 − λ ′ 1 − ˆψψ 2 , (7.168)<br />

see Eq. (7.160). Let us have a look at the classical field equations, namely<br />

δA/δψ = 0, which is always solved by ˆψ = 1, reflect<strong>in</strong>g probability conservation,<br />

and δA/δ ˆψ = 0, which, upon <strong>in</strong>sert<strong>in</strong>g ˆψ = 1 gives here<br />

∂ψ(x,t)<br />

= D ∇ 2 ψ(x,t) − (2λ + λ ′ ) ψ(x,t) 2 , (7.169)<br />

∂t<br />

i.e., essentially the mean-field rate equation for the local particle density<br />

ψ(x,t), see Eq. (7.149), supplemented with diffusion. The field theory action<br />

(7.167), derived from the master equation (7.154), then provides a means<br />

of <strong>in</strong>clud<strong>in</strong>g fluctuations <strong>in</strong> our analysis.<br />

Before we proceed with this program, it is <strong>in</strong>structive to perform a shift<br />

<strong>in</strong> the field ˆψ about the mean-field solution, ˆψ(x, t) =1+˜ψ(x, t), whereupon<br />

the reaction Hamiltonian density (7.168) becomes<br />

H reac ( ˜ψ, ψ) =(2λ + λ ′ ) ˜ψψ 2 +(λ + λ ′ ) ˜ψ 2 ψ 2 . (7.170)<br />

In addition to the diffusion propagator, the annihilation and coagulation<br />

processes thus give identical three- and four-po<strong>in</strong>t vertices; aside from nonuniversal<br />

amplitudes, one should therefore obta<strong>in</strong> identical scal<strong>in</strong>g behaviour<br />

for both b<strong>in</strong>ary reactions <strong>in</strong> the asymptotic long-time limit [57]. Lastly, we<br />

remark that if we <strong>in</strong>terpret the action A[ ˜ψ, ψ] as a response functional (7.34),<br />

despite the fields ˜ψ not be<strong>in</strong>g purely imag<strong>in</strong>ary, our field theory becomes formally<br />

equivalent to a “Langev<strong>in</strong>” equation, where<strong>in</strong> additive noise is added to


334 U. C. Täuber<br />

Eq. (7.169), albeit with negative correlator L[ψ] =−(λ + λ ′ ) ψ 2 , which represents<br />

“imag<strong>in</strong>ary” multiplicative noise. This Langev<strong>in</strong> description is thus not<br />

well-def<strong>in</strong>ed; however, one may render the noise correlator positive through<br />

a nonl<strong>in</strong>ear Cole–Hopf transformation ˜ψ =e˜ρ , ψ = ρ e −˜ρ such that ˜ψψ = ρ,<br />

with Jacobian 1, but at the expense of “diffusion noise” ∝ Dρ(∇˜ρ ) 2 <strong>in</strong> the action<br />

[58]. In summary, b<strong>in</strong>ary (and higher-order) annihilation and coagulation<br />

processes cannot be cast <strong>in</strong>to a Langev<strong>in</strong> framework <strong>in</strong> any simple manner.<br />

7.2.3 Diffusion-Limited S<strong>in</strong>gle-Species Annihilation Processes<br />

We beg<strong>in</strong> by analys<strong>in</strong>g diffusion-limited s<strong>in</strong>gle-species annihilation kA →∅<br />

[57, 59]. The correspond<strong>in</strong>g field theory action (7.167) reads<br />

∫ ∫ [ ( )<br />

A[ ˆψ, ∂<br />

ψ] = d d x dt ˆψ<br />

∂t − D∇2 ψ − λ<br />

(1 − ˆψ k) ]<br />

ψ k , (7.171)<br />

which for k ≥ 3 allows no (obvious) equivalent Langev<strong>in</strong> description. Straightforward<br />

power count<strong>in</strong>g gives the scal<strong>in</strong>g dimension for the annihilation rate,<br />

[λ] =µ 2−(k−1)d , which suggests the upper critical dimension d c (k) =2/(k−1).<br />

Thus we expect mean-field behaviour ∼ (λt) −1/(k−1) , see Eq. (7.151), <strong>in</strong> any<br />

physical dimension for k>3, logarithmic corrections at d c =1fork =3and<br />

at d c =2fork = 2, and nonclassical power laws for pair annihilation only <strong>in</strong><br />

one dimension. The field theory def<strong>in</strong>ed by the action (7.171) has two vertices,<br />

the “annihilation” s<strong>in</strong>k with k <strong>in</strong>com<strong>in</strong>g l<strong>in</strong>es only, and the “scatter<strong>in</strong>g” vertex<br />

with k <strong>in</strong>com<strong>in</strong>g and k outgo<strong>in</strong>g l<strong>in</strong>es. Neither allows for propagator renormalisation,<br />

hence the model rema<strong>in</strong>s massless with exact scal<strong>in</strong>g exponents<br />

η =0andz = 2, i.e., diffusive dynamics.<br />

In addition, the entire perturbation expansion for the renormalisation for<br />

the annihilation vertices is merely a geometric series of the one-loop diagram,<br />

see Fig. 7.6 for the pair annihilation case (k = 2). If we def<strong>in</strong>e the renormalised<br />

effective coupl<strong>in</strong>g accord<strong>in</strong>g to<br />

g R = Z g<br />

λ<br />

D B kd µ −2(1−d/dc) , (7.172)<br />

where B kd = k! Γ (2−d/d c ) d c /k d/2 (4π) d/dc , we obta<strong>in</strong> for the s<strong>in</strong>gle nontrivial<br />

renormalisation constant<br />

+ + +<br />

. . .<br />

+<br />

. . .<br />

+<br />

. . .<br />

Fig. 7.6. Vertex renormalisation for diffusion-limited b<strong>in</strong>ary annihilation A+A →∅


7 Field-Theory Approaches to Nonequilibrium Dynamics 335<br />

Z −1<br />

g<br />

=1+ λB kd µ −2(1−d/dc)<br />

D (d c − d)<br />

(7.173)<br />

to all orders. Consequently, the associated RG beta function becomes<br />

β g = µ ∂<br />

∂µ ∣ g R = − 2 g R<br />

(d − d c + g R ) , (7.174)<br />

0<br />

d c<br />

with the Gaussian fixed po<strong>in</strong>t g ∗ 0 = 0 stable for d > d c (k) = 2/(k − 1),<br />

lead<strong>in</strong>g to the mean-field power laws (7.151), whereas for d


336 U. C. Täuber<br />

difference of particle numbers (even locally), i.e., there is a conservation law<br />

for c(t) =a(t) − b(t) =c(0) [61]. The rate equations for the concentrations<br />

ȧ(t) =−λa(t) b(t) =ḃ(t) (7.181)<br />

are for equal <strong>in</strong>itial densities a(0) = b(0) solved by the s<strong>in</strong>gle-species pair<br />

annihilation mean-field power law<br />

a(t) =b(t) ∼ (λt) −1 , (7.182)<br />

whereas for unequal <strong>in</strong>itial densities c(0) = a(0) − b(0) > 0, say, the majority<br />

species a(t) → a ∞ = c(0) > 0ast →∞, and the m<strong>in</strong>ority population disappears,<br />

b(t) → 0. From Eq. (7.181) we obta<strong>in</strong> for d>d c = 2 the exponential<br />

approach<br />

a(t) − a ∞ ∼ b(t) ∼ e −c(0) λt . (7.183)<br />

Mapp<strong>in</strong>g the associated master equation onto a cont<strong>in</strong>uum field theory<br />

(7.167), the reaction term now reads (with the fields ψ and ϕ represent<strong>in</strong>g the<br />

A and B particles, respectively) [62]<br />

H reac ( ˆψ,<br />

(<br />

ψ, ˆϕ, ϕ) =−λ 1 − ˆψ<br />

)<br />

ˆϕ ψϕ . (7.184)<br />

As <strong>in</strong> the s<strong>in</strong>gle-species case, there is no propagator renormalisation, and<br />

moreover the Feynman diagrams for the renormalised reaction vertex are of<br />

precisely the same form as for A + A →∅, see Fig. 7.6. Thus, for unequal<br />

<strong>in</strong>itial densities, c(0) > 0, the mean-field power law ∼ λt <strong>in</strong> the exponent of<br />

Eq. (7.183) becomes aga<strong>in</strong> replaced with (Dt) d/2 <strong>in</strong> dimensions d ≤ d c =2,<br />

lead<strong>in</strong>g to stretched exponential time dependence,<br />

d


7 Field-Theory Approaches to Nonequilibrium Dynamics 337<br />

∫<br />

c(x,t)= d d x ′ G 0 (x − x ′ ,t) c(x ′ , 0) . (7.188)<br />

Let us furthermore assume a Poisson distribution for the <strong>in</strong>itial density correlations<br />

(<strong>in</strong>dicated by an overbar), a(x, 0) a(x ′ , 0) = a(0) 2 + a(0) δ(x − x ′ )=<br />

b(x, 0) b(x ′ , 0) and a(x, 0) b(x ′ (0) = a(0) 2 , which implies c(x, 0) c(x ′ , 0) =<br />

2 a(0)δ(x−x ′ ). Averag<strong>in</strong>g over the <strong>in</strong>itial conditions then yields with Eq. (7.188)<br />

∫<br />

c(x,t) 2 =2a(0) d d x ′ G 0 (x − x ′ ,t) 2 =2a(0) (8πDt) −d/2 ; (7.189)<br />

s<strong>in</strong>ce the distribution for c will be a Gaussian, we thus obta<strong>in</strong> for the local<br />

density excess orig<strong>in</strong>at<strong>in</strong>g <strong>in</strong> a random <strong>in</strong>itial fluctuation,<br />

√ √<br />

2 a(0)<br />

|c(x,t)| =<br />

π c(x,t)2 =2<br />

π (8πDt)−d/4 . (7.190)<br />

In dimensions d2and∼t −d/2 for d


338 U. C. Täuber<br />

and D + A →∅. We may then obviously identify A = C and B = D, which<br />

leads back to the case of two-species pair annihilation. Generally, with<strong>in</strong> essentially<br />

mean-field theory one f<strong>in</strong>ds for cyclic multi-species annihilation processes<br />

a i (t) ∼ t −α(q,d) , where for<br />

{<br />

4 q =2, 4, 6,...<br />

2


7 Field-Theory Approaches to Nonequilibrium Dynamics 339<br />

[λ] =µ 2−d ,isthusirrelevant <strong>in</strong> the RG sense, and can be dropped for the<br />

computation of universal, asymptotic scal<strong>in</strong>g properties.<br />

The action (7.195) with λ =0isknownasReggeon field theory [65], and<br />

its basic characteristic is its <strong>in</strong>variance under rapidity <strong>in</strong>version S(x,t) ↔<br />

− ˜S(x, −t). If we <strong>in</strong>terpret Eq. (7.195) as a response functional, we see that it<br />

becomes formally equivalent to a stochastic process with multiplicative noise<br />

(〈ζ(x,t)〉 = 0) captured by the Langev<strong>in</strong> equation [66, 67]<br />

∂S(x,t)<br />

= −D ( r − ∇ 2) S(x,t) − uS(x,t) 2 + ζ(x,t) ,<br />

∂t<br />

(7.196)<br />

〈ζ(x,t) ζ(x ′ ,t ′ )〉 =2uS(x,t) δ(x − x ′ ) δ(t − t ′ ) (7.197)<br />

(for a more accurate mapp<strong>in</strong>g procedure, see Ref. [68]), which is essentially a<br />

noisy Fisher–Kolmogorov equation (7.193), with the noise correlator (7.197)<br />

ensur<strong>in</strong>g that the fluctuations <strong>in</strong>deed cease <strong>in</strong> the absorb<strong>in</strong>g state where<br />

〈S〉 = 0. It has moreover been established [69–71] that the action (7.195) describes<br />

the scal<strong>in</strong>g properties of critical directed percolation (DP) clusters [72],<br />

illustrated <strong>in</strong> Fig. 7.7. Indeed, if the DP growth direction is labeled as “time”<br />

t, we see that the structure of the DP clusters emerges from the basic decay,<br />

branch<strong>in</strong>g, and coagulation reactions encoded <strong>in</strong> Eq. (7.194).<br />

t ⃗<br />

Fig. 7.7. Directed percolation process (left) and critical DP cluster (right)<br />

In fact, the field theory action should govern the scal<strong>in</strong>g properties of<br />

generic cont<strong>in</strong>uous nonequilibrium phase transitions from active to <strong>in</strong>active,<br />

absorb<strong>in</strong>g states, namely for an order parameter with Markovian stochastic<br />

dynamics that is decoupled from any other slow variable, and <strong>in</strong> the absence<br />

of quenched randomness [71, 73]. This DP conjecture follows from the follow<strong>in</strong>g<br />

phenomenological approach [68] to simple epidemic processes (SEP), or<br />

epidemics with recovery [46]:<br />

1. A “susceptible” medium becomes locally “<strong>in</strong>fected”, depend<strong>in</strong>g on the density<br />

n of neighbor<strong>in</strong>g “sick” <strong>in</strong>dividuals. The <strong>in</strong>fected regions recover after<br />

a brief time <strong>in</strong>terval.<br />

2. The state n =0isabsorb<strong>in</strong>g. It describes the ext<strong>in</strong>ction of the “disease’.<br />

3. The disease spreads out diffusively via the short-range <strong>in</strong>fection 1. of neighbor<strong>in</strong>g<br />

susceptible regions.


340 U. C. Täuber<br />

4. Microscopic fast degrees of freedom are <strong>in</strong>corporated as local noise or<br />

stochastic forces that respect statement 2., i.e., the noise alone cannot<br />

regenerate the disease.<br />

These <strong>in</strong>gredients are captured by the coarse-gra<strong>in</strong>ed mesoscopic Langev<strong>in</strong><br />

equation ∂ t n = D ( ∇ 2 − R[n] ) n + ζ with a reaction functional R[n], and<br />

s stochastic noise correlator of the form L[n] =nN[n]. Near the ext<strong>in</strong>ction<br />

threshold, we may expand R[n] =r + un + ..., N[n] =v + ..., and higherorder<br />

terms turn out to be irrelevant <strong>in</strong> the RG sense. Upon rescal<strong>in</strong>g, we<br />

recover the Reggeon field theory action (7.195) for DP as the correspond<strong>in</strong>g<br />

response functional (7.34).<br />

We now proceed to an explicit evaluation of the DP critical exponents<br />

to one-loop order, closely follow<strong>in</strong>g the recipes given <strong>in</strong> Subsects. 7.1.3–7.1.5.<br />

The lowest-order fluctuation contribution to the two-po<strong>in</strong>t vertex function<br />

Γ (1,1) (q,ω) (propagator self-energy) is depicted <strong>in</strong> Fig. 7.8(a). The Feynman<br />

rules of Subsect. 7.1.3 yield the correspond<strong>in</strong>g analytic expression<br />

Γ (1,1) (q,ω)=iω + D (r + q 2 )+ u2<br />

D<br />

∫<br />

k<br />

1<br />

iω/2D + r + q 2 /4+k 2 . (7.198)<br />

The criticality condition Γ (1,1) (0, 0) = 0 at r = r c<br />

fluctuation-<strong>in</strong>duced shift of the percolation threshold<br />

provides us with the<br />

∫ Λ<br />

r c = − u2 1<br />

D 2 r c + k 2 + O(u4 ) . (7.199)<br />

k<br />

Insert<strong>in</strong>g τ = r − r c <strong>in</strong>to Eq. (7.198), we then f<strong>in</strong>d to this order<br />

∫<br />

Γ (1,1) (q,ω)=iω + D (τ + q 2 ) − u2 iω/2D + τ + q 2 /4<br />

D k 2 (iω/2D + τ + q 2 /4+k 2 ) , (7.200)<br />

k<br />

and the diagram <strong>in</strong> Fig. 7.8(b) for the three-po<strong>in</strong>t vertex functions, evaluated<br />

at zero external wavevectors and frequencies, gives<br />

(<br />

)<br />

Γ (1,2) ({0}) =−Γ (2,1) ({0}) =−2u 1 − 2u2 1<br />

D<br />

∫k<br />

2 (τ + k 2 ) 2 . (7.201)<br />

For the renormalisation factors, we use aga<strong>in</strong> the conventions (7.56) and<br />

(7.58), but with<br />

(a)<br />

(b)<br />

Fig. 7.8. DP renormalisation: one-loop diagrams for the vertex functions (a) Γ (1,1)<br />

(propagator self-energy), and (b) Γ (1,2) = −Γ (2,1) (nonl<strong>in</strong>ear vertices)


7 Field-Theory Approaches to Nonequilibrium Dynamics 341<br />

u R = Z u uA 1/2<br />

d<br />

µ (d−4)/2 . (7.202)<br />

Because of rapidity <strong>in</strong>version <strong>in</strong>variance, Z ˜S<br />

= Z S . With Eq. (7.53) the derivatives<br />

of Γ (1,1) with respect to ω, q 2 ,andτ, as well as the one-loop result for<br />

Γ (1,2) <strong>in</strong> Eq. (7.201), all evaluated at the normalisation po<strong>in</strong>t τ R =1,provide<br />

us with the Z factors<br />

Z S =1− u2 A d µ −ɛ<br />

2D 2 ɛ<br />

Z τ =1− 3u2 A d µ −ɛ<br />

4D 2 ɛ<br />

From these we <strong>in</strong>fer the RG flow functions<br />

, Z D =1+ u2 A d µ −ɛ<br />

4D 2 ɛ<br />

, Z u =1− 5u2 A d µ −ɛ<br />

4D 2 ɛ<br />

,<br />

. (7.203)<br />

γ S = v R /2 , γ D = −v R /4 , γ τ = −2+3v R /4 , (7.204)<br />

with the renormalised effective coupl<strong>in</strong>g<br />

v R = Z2 u<br />

Z 2 D<br />

whose RG beta function is to this order<br />

u 2<br />

D 2 A d µ d−4 , (7.205)<br />

β v = v R<br />

[<br />

−ɛ +3vR + O(v 2 R) ] . (7.206)<br />

For d>d c = 4, the Gaussian fixed po<strong>in</strong>t v ∗ 0 = 0 is stable, and we recover<br />

the mean-field critical exponents. For ɛ =4− d>0, we f<strong>in</strong>d the nontrivial<br />

IR-stable RG fixed po<strong>in</strong>t<br />

v ∗ = ɛ/3+O(ɛ 2 ) . (7.207)<br />

Sett<strong>in</strong>g up and solv<strong>in</strong>g the RG equation (7.65) for the vertex function<br />

proceeds just as <strong>in</strong> Subsect. 7.1.5. With the identifications (7.83) (with a =0)<br />

we thus obta<strong>in</strong> the DP critical exponents to first order <strong>in</strong> ɛ,<br />

η = − ɛ 6 + O(ɛ2 ) ,<br />

1<br />

ν =2− ɛ 4 + O(ɛ2 ) ,<br />

z =2− ɛ<br />

12 + O(ɛ2 ) . (7.208)<br />

In the vic<strong>in</strong>ity of v ∗ , the solution of the RG equation for the order parameter<br />

reads, recall<strong>in</strong>g that [S] =µ d/2 ,<br />

)<br />

〈S R (τ R ,t)〉 ≈µ d/2 l (d−γ∗ S )/2 Ŝ<br />

(τ R l γ∗ τ ,v<br />

∗<br />

R ,D R µ 2 l 2+γ∗ D t , (7.209)<br />

which leads to the follow<strong>in</strong>g scal<strong>in</strong>g relations and explicit exponent values,<br />

β =<br />

ν(d + η)<br />

2<br />

=1− ɛ 6 + O(ɛ2 ) , α = β<br />

zν =1− ɛ 4 + O(ɛ2 ) . (7.210)<br />

The scal<strong>in</strong>g exponents for critical directed percolation are known analytically<br />

for a plethora of physical quantities (but the reader should beware that


342 U. C. Täuber<br />

various different conventions are used <strong>in</strong> the literature); for the two-loop results<br />

to order ɛ 2 <strong>in</strong> the perturbative dimensional expansion, see Ref. [68]. In<br />

Table 7.2, we compare the O(ɛ) values with the results from Monte Carlo<br />

computer simulations, which allow the DP critical exponents to be measured<br />

to high precision (for recent overviews on simulation results for DP and other<br />

absorb<strong>in</strong>g state phase transitions, see Refs. [74, 75]). Yet unfortunately, there<br />

are to date hardly any real experiments that would confirm the DP conjecture<br />

[71, 73] and actually measure the scal<strong>in</strong>g exponents for this prom<strong>in</strong>ent<br />

nonequilibrium universality class.<br />

Table 7.2. Comparison of the DP critical exponent values from Monte Carlo simulations<br />

with the results from the ɛ expansion<br />

Scal<strong>in</strong>g exponent d =1 d =2 d =4− ɛ<br />

ξ ∼|τ| −ν ν ≈ 1.100 ν ≈ 0.735 ν =1/2+ɛ/16 + O(ɛ 2 )<br />

t c ∼ ξ z ∼|τ| −zν z ≈ 1.576 z ≈ 1.73 z =2− ɛ/12 + O(ɛ 2 )<br />

a ∞ ∼|τ| β β ≈ 0.2765 β ≈ 0.584 β =1− ɛ/6+O(ɛ 2 )<br />

a c(t) ∼ t −α α ≈ 0.160 α ≈ 0.46 α =1− ɛ/4+O(ɛ 2 )<br />

7.2.6 Dynamic Isotropic Percolation and Multi-Species Variants<br />

An <strong>in</strong>terest<strong>in</strong>g variant of active to absorb<strong>in</strong>g state phase transitions emerges<br />

when we modify the SEP rules (1) and (2) <strong>in</strong> Subsect. 7.2.5 to<br />

1. The susceptible medium becomes <strong>in</strong>fected, depend<strong>in</strong>g on the densities n<br />

and m of sick <strong>in</strong>dividuals and the “debris”, respectively. After a brief time<br />

<strong>in</strong>terval, the sick <strong>in</strong>dividuals decay <strong>in</strong>to immune debris, which ultimately<br />

stops the disease locally by exhaust<strong>in</strong>g the supply of susceptible regions.<br />

2. The states with n = 0 and any spatial distribution of m are absorb<strong>in</strong>g,<br />

and describe the ext<strong>in</strong>ction of the disease.<br />

Here, the debris is given by the accumulated decay products,<br />

∫ t<br />

m(x,t)=κ n(x,t ′ )dt ′ . (7.211)<br />

−∞<br />

After rescal<strong>in</strong>g, this general epidemic process (GEP) or epidemic with removal<br />

[46] is described <strong>in</strong> terms of the mesoscopic Langev<strong>in</strong> equation [76]<br />

∂S(x,t)<br />

= −D ( r − ∇ 2) ∫ t<br />

S(x,t) − DuS(x,t) S(x,t ′ ) Dt ′ + ζ(x,t) ,<br />

∂t<br />

−∞<br />

(7.212)<br />

with noise correlator (7.197). The associated response functional reads [77,78]


7 Field-Theory Approaches to Nonequilibrium Dynamics 343<br />

∫ ∫ [ (<br />

A[ ˜S,S]=<br />

∂<br />

d d x dt ˜S<br />

∂t + D ( r − ∇ 2)) t<br />

]<br />

S − u ˜S 2 S + DuS∫<br />

S(t ′ ) .<br />

(7.213)<br />

For the field theory thus def<strong>in</strong>ed, one may take the quasistatic limit by<br />

<strong>in</strong>troduc<strong>in</strong>g the fields<br />

˜ϕ(x) = ˜S(x,t→∞) , ϕ(x) =D<br />

∫ ∞<br />

−∞<br />

S(x,t ′ )dt ′ . (7.214)<br />

For t →∞, the action (7.213) thus becomes<br />

∫ [<br />

]<br />

A qst [˜ϕ, ϕ] = d d x ˜ϕ r − ∇ 2 − u (˜ϕ − ϕ) ϕ, (7.215)<br />

which is known to describe the critical exponents of isotropic percolation [79].<br />

An isotropic percolation cluster is shown <strong>in</strong> Fig. 7.9, to be contrasted with<br />

the anisotropic scal<strong>in</strong>g evident <strong>in</strong> Fig. 7.7(b). The upper critical dimension of<br />

isotropic percolation is d c = 6, and an explicit calculation, with the diagrams<br />

of Fig. 7.8, but <strong>in</strong>volv<strong>in</strong>g the static propagators G 0 (q) =1/(r +q 2 ), yields the<br />

follow<strong>in</strong>g critical exponents for isotropic percolation, to first order <strong>in</strong> ɛ =6−d,<br />

η = − ɛ<br />

21 + O(ɛ2 ) ,<br />

1<br />

ν =2− 5 ɛ<br />

21 + O(ɛ2 ) , β =1− ɛ 7 + O(ɛ2 ) . (7.216)<br />

In order to calculate the dynamic critical exponent for this dynamic isotropic<br />

percolation (dIP) universality class, we must return to the full action (7.213).<br />

Once aga<strong>in</strong>, with the diagrams of Fig. 7.8, but now <strong>in</strong>volv<strong>in</strong>g a temporally<br />

nonlocal three-po<strong>in</strong>t vertex, one then arrives at<br />

z =2− ɛ 6 + O(ɛ2 ) . (7.217)<br />

For a variety of two-loop results, the reader is referred to Ref. [68]. It is also<br />

possible to describe the crossover from isotropic to directed percolation with<strong>in</strong><br />

this field-theoretic framework [80, 81].<br />

Fig. 7.9. Isotropic percolation cluster


344 U. C. Täuber<br />

Let us next consider multi-species variants of directed percolation processes,<br />

which can be obta<strong>in</strong>ed <strong>in</strong> the particle language by coupl<strong>in</strong>g the DP reactions<br />

A i →∅, A i ⇋ A i + A i via processes of the form A i ⇋ A j + A j (with j ≠ i);<br />

or directly by the correspond<strong>in</strong>g generalisation with<strong>in</strong> the Langev<strong>in</strong> representation<br />

with 〈ζ i (x,t)〉 =0,<br />

∂S i<br />

∂t = D (<br />

i ∇ 2 − R i [S i ] ) S i + ζ i , R i [S i ]=r i + ∑ j<br />

g ij S j + ... , (7.218)<br />

〈ζ i (x,t)ζ j (x ′ ,t ′ )〉=2S i N i [S i ] δ(x − x ′ ) δ(t − t ′ ) δ ij ,N i [S i ]=u i + ... (7.219)<br />

The ensu<strong>in</strong>g renormalisation factors turn out to be precisely as for s<strong>in</strong>glespecies<br />

DP, and consequently the generical critical behaviour even <strong>in</strong> such<br />

multi-species systems is governed by the DP universality class [58]. For example,<br />

the predator ext<strong>in</strong>ction threshold for the stochastic Lotka–Volterra<br />

system mentioned <strong>in</strong> Subsect. 7.2.1 is characterised by the DP exponents as<br />

well [47]. But these reactions also generate A i → A j , caus<strong>in</strong>g additional terms<br />

∑<br />

j≠i g j S j <strong>in</strong> Eq. (7.218). Asymptotically, the <strong>in</strong>ter-species coupl<strong>in</strong>gs become<br />

unidirectional, which allows for the appearance of special multicritical po<strong>in</strong>ts<br />

when several r i = 0 simultaneously [82]. This leads to a hierarchy of order<br />

parameter exponents β k on the kth level of a unidirectional cascade, with<br />

β 1 =1− ɛ 6 + O(ɛ2 ) ,<br />

β 2 = 1 2 − 13 ɛ<br />

96 + O(ɛ2 ) ,..., β k = 1 − O(ɛ) ; (7.220)<br />

2k for the associated crossover exponent, one can show Φ =1toall orders [58].<br />

Quite analogous features emerge for multi-species dIP processes [58, 68].<br />

7.2.7 Conclud<strong>in</strong>g Remarks<br />

In these lecture notes, I have described how stochastic processes can be<br />

mapped onto field theory representations, start<strong>in</strong>g either from a mesoscopic<br />

Langev<strong>in</strong> equation for the coarse-gra<strong>in</strong>ed densities of the relevant order parameter<br />

fields and conserved quantities, or from a more microsopic master<br />

equation for <strong>in</strong>teract<strong>in</strong>g particle systems. The dynamic renormalisation group<br />

method can then be employed to study and characterise the universal scal<strong>in</strong>g<br />

behaviour near cont<strong>in</strong>uous phase transitions both <strong>in</strong> and far from thermal<br />

equilibrium, and for systems that generically display scale-<strong>in</strong>variant features.<br />

While the critical dynamics near equilibrium phase transitions has been thoroughly<br />

<strong>in</strong>vestigated experimentally <strong>in</strong> the past three decades, regrettably such<br />

direct experimental verification of the by now considerable amount of theoretical<br />

work on nonequilibrium systems is largely amiss. In this respect, applications<br />

of the expertise ga<strong>in</strong>ed <strong>in</strong> the nonequilibrium statistical mechanics<br />

of complex cooperative behaviour to biological systems might prove fruitful<br />

and constitutes a promis<strong>in</strong>g venture. One must bear <strong>in</strong> m<strong>in</strong>d, however, that<br />

nonuniversal features are often crucial for the relevant questions <strong>in</strong> biology.


7 Field-Theory Approaches to Nonequilibrium Dynamics 345<br />

In part, the lack of clearcut experimental evidence may be due to the fact<br />

that asymptotic universal properties are perhaps less prom<strong>in</strong>ent <strong>in</strong> accessible<br />

nonequilibrium systems, ow<strong>in</strong>g to long crossover times. Yet fluctuations do<br />

tend to play a more important role <strong>in</strong> systems that are driven away from thermal<br />

equilibrium, and the concept of universality classes, despite the undoubtedly<br />

much <strong>in</strong>creased richness <strong>in</strong> dynamical systems, should still be useful. For<br />

example, we have seen that the directed percolation universality class quite<br />

generically describes the critical properties of phase transitions from active<br />

to <strong>in</strong>active, absorb<strong>in</strong>g states, which abound <strong>in</strong> nature. The few exceptions<br />

to this rule either require the coupl<strong>in</strong>g to another conserved mode [83, 84];<br />

the presence, on a mesoscopic level, of additional symmetries that preclude<br />

the spontaneous decay A → ∅ as <strong>in</strong> the so-called parity-conserv<strong>in</strong>g (PC)<br />

universality class, represented by branch<strong>in</strong>g and annihilat<strong>in</strong>g random walks<br />

A → (n +1)A with n even, andA + A →∅[85] (for recent developments<br />

based on nonperturbative RG approaches, see Ref. [86]); or the absence of any<br />

first-order reactions, as <strong>in</strong> the (by now rather notorious) pair contact process<br />

with diffusion (PCPD) [87], which has so far eluded a successful field-theoretic<br />

treatment [88]. A possible explanation for the fact that DP exponents have not<br />

been measured ubiquitously (yet) could be the <strong>in</strong>stability towards quenched<br />

disorder <strong>in</strong> the reaction rates [89].<br />

In reaction–diffusion systems, a complete classification of the scal<strong>in</strong>g properties<br />

<strong>in</strong> multi-species systems rema<strong>in</strong>s <strong>in</strong>complete, aside from pair annihilation<br />

and DP-like processes, and still constitutes a quite formidable program<br />

(for a recent overview over the present situation from a field-theoretic viewpo<strong>in</strong>t,<br />

see Ref. [12]). This is even more evident for nonequilibrium systems <strong>in</strong><br />

general, even when ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> driven steady states. Field-theoretic methods<br />

and the dynamic renormalisation group represent powerful tools that I<br />

believe will cont<strong>in</strong>ue to crucially complement exact solutions (usually of onedimensional<br />

models), other approximative approaches, and computer simulations,<br />

<strong>in</strong> our quest to further elucidate the <strong>in</strong>trigu<strong>in</strong>g cooperative behaviour<br />

of strongly <strong>in</strong>teract<strong>in</strong>g and fluctuat<strong>in</strong>g many-particle systems.<br />

Acknowledgements<br />

This work has been supported <strong>in</strong> part by the U.S. National Science Foundation<br />

through grant NSF DMR-0308548. I am <strong>in</strong>debted to many colleagues,<br />

students, and friends, from and with whom I had the pleasure to learn and<br />

research the material presented here; specifically I would like to mention<br />

Vamsi Akk<strong>in</strong>eni, Michael Bulenda, John Cardy, Olivier Deloubrière, Daniel<br />

Fisher, Re<strong>in</strong>hard Folk, Erw<strong>in</strong> Frey, Ivan Georgiev, Yad<strong>in</strong> Goldschmidt, Peter<br />

Grassberger, Henk Hilhorst, Haye H<strong>in</strong>richsen, Mart<strong>in</strong> Howard, Terry Hwa,<br />

Hannes Janssen, Bernhard Kaufmann, Mauro Mobilia, David Nelson, Beth<br />

Reid, Zoltán Rácz, Jaime Santos, Beate Schmittmann, Franz Schwabl, Steffen<br />

Trimper, Ben Vollmayr-Lee, Mark Washenberger, Fred van Wijland, and<br />

Royce Zia. Lastly, I would like to thank the organisers of the very enjoyable


346 U. C. Täuber<br />

Luxembourg Summer School on Age<strong>in</strong>g and the Glass Transition for their<br />

k<strong>in</strong>d <strong>in</strong>vitation, and my colleagues at the Laboratoire de Physique Théorique,<br />

Université de Paris-Sud Orsay, France and at the Rudolf Peierls Centre for<br />

Theoretical <strong>Physics</strong>, University of Oxford, U.K., where these lecture notes<br />

were conceived and written, for their warm hospitality.<br />

References<br />

1. P. Ramond: Field theory – a modern primer, (Benjam<strong>in</strong>/Cumm<strong>in</strong>gs, Read<strong>in</strong>g<br />

1981)<br />

2. D.J. Amit: Field theory, the renormalization group, and critical phenomena<br />

(World Scientific, S<strong>in</strong>gapore 1984)<br />

3. C. Itzykson and J.M. Drouffe: Statistical field theory (Cambridge University<br />

Press, Cambridge 1989)<br />

4. M. Le Bellac: Quantum and statistical field theory, (Oxford University Press,<br />

Oxford 1991)<br />

5. J. Z<strong>in</strong>n-Just<strong>in</strong>: Quantum field theory and critical phenomena (Clarendon Press,<br />

Oxford 1993)<br />

6. J. Cardy: Scal<strong>in</strong>g and renormalization <strong>in</strong> statistical physics (Cambridge University<br />

Press, Cambridge 1996)<br />

7. P.C. Hohenberg and B.I. Halper<strong>in</strong>: Rev. Mod. Phys. 49, 435 (1977)<br />

8. H.K. Janssen: Field-theoretic methods applied to critical dynamics. In: Dynamical<br />

critical phenomena and related topics, <strong>Lecture</strong> <strong>Notes</strong> <strong>in</strong> <strong>Physics</strong>, vol. 104,<br />

ed by C.P. Enz (<strong>Spr<strong>in</strong>ger</strong>, Heidelberg 1979), pp. 26–47<br />

9. R. Bausch, H.K. Janssen, and H. Wagner: Z. Phys. B24, 113 (1976)<br />

10. J.L. Cardy: Renormalisation group approach to reaction-diffusion problems. In:<br />

Proceed<strong>in</strong>gs of Mathematical Beauty of <strong>Physics</strong>, edbyJ.-B.Zuber,Adv.Ser.<br />

<strong>in</strong> Math. Phys. 24, 113 (1997)<br />

11. D.C. Mattis and M.L. Glasser: Rev. Mod. Phys. 70, 979 (1998)<br />

12. U.C. Täuber, M.J. Howard, and B.P. Vollmayr-Lee: J. Phys. A: Math. Gen.<br />

38, R79 (2005)<br />

13. U.C. Täuber: Critical dynamics: a field theory approach to equilibrium and<br />

non-equilibrium scal<strong>in</strong>g behavior, <strong>in</strong> preparation (to be published at Cambridge<br />

University Press, Cambridge); for completed chapters, see:<br />

http://www.phys.vt.edu/˜tauber/utaeuber.html<br />

14. R.A. Ferrell, N. Menyhàrd, H. Schmidt, F. Schwabl, and P. Szépfalusy: Phys.<br />

Rev. Lett. 18, 891 (1967); Ann. of Phys. 47, 565 (1968)<br />

15. B.I. Halper<strong>in</strong> and P.C. Hohenberg: Phys. Rev. 177, 952 (1969)<br />

16. B.I. Halper<strong>in</strong>, P.C. Hohenberg, and S.-k. Ma: Phys. Rev. Lett. 29, 1548 (1972)<br />

17. C. De Dom<strong>in</strong>icis, E. Bréz<strong>in</strong>, and J. Z<strong>in</strong>n-Just<strong>in</strong>: Phys. Rev. B12, 4945 (1975)<br />

18. H.K. Janssen: Z. Phys. B23, 377 (1976)<br />

19. C. De Dom<strong>in</strong>icis: J. Physique Colloque 37, C2247 (1976)<br />

20. P.C. Mart<strong>in</strong>, E.D. Siggia, and H.A. Rose: Phys. Rev. A8, 423 (1973)<br />

21. H. Wagner: Z. Phys. 195, 273 (1966)<br />

22. N.D. Merm<strong>in</strong> and H. Wagner: Phys. Rev. Lett. 17, 1133 (1966)<br />

23. P.C. Hohenberg: Phys. Rev. 158, 383 (1967)<br />

24. P.M. Chaik<strong>in</strong> and T.C. Lubensky: Pr<strong>in</strong>ciples of condensed matter physics,<br />

(Cambridge University Press, Cambridge 1995)


7 Field-Theory Approaches to Nonequilibrium Dynamics 347<br />

25. S.-k. Ma and G.F. Mazenko: Phys. Rev. Lett. 33, 1383 (1974); Phys. Rev. B11,<br />

4077 (1975)<br />

26. E. Frey and F. Schwabl: Adv. Phys. 43, 577 (1994)<br />

27. H.K. Janssen, B. Schaub, and B. Schmittmann: Z. Phys. B73, 539 (1989)<br />

28. H.K. Janssen: On the renormalized field theory of nonl<strong>in</strong>ear critical relaxation.<br />

In: From phase transitions to chaos, edbyG.Györgyi, I. Kondor, L. Sasvári,<br />

and T. Tél (World Scientific, S<strong>in</strong>gapore 1992), pp. 68–91.<br />

29. P. Calabrese and A. Gambassi: Phys. Rev. E66, 066101 (2002); J. Phys. A:<br />

Math. Gen. 38, R133 (2005)<br />

30. A. Gambassi: In: Proceed<strong>in</strong>gs of the International Summer School “Age<strong>in</strong>g and<br />

the Glass Transition”, to appear <strong>in</strong> J. Phys. Conf. Proc. (July 2006)<br />

31. H.W. Diehl: In: Phase Transitions and Critical Phenomena, vol. 10, ed by<br />

C. Domb and J.L. Lebowitz (Academic Press, London 1986)<br />

32. K. Oerd<strong>in</strong>g and H.K. Janssen: J. Phys. A: Math. Gen. 26, 5295 (1993)<br />

33. U.C. Täuber, V.K. Akk<strong>in</strong>eni, and J.E. Santos: Phys. Rev. Lett. 88, 045702<br />

(2002)<br />

34. F. Haake, M. Lewenste<strong>in</strong>, and M. Wilkens: Z. Phys. B55, 211 (1984)<br />

35. G. Gr<strong>in</strong>ste<strong>in</strong>, C. Jayaprakash, and Y. He: Phys. Rev. Lett. 55, 2527 (1985)<br />

36. K.E. Bassler and B. Schmittmann: Phys. Rev. Lett. 73, 3343 (1994)<br />

37. U.C. Täuber and Z. Rácz: Phys. Rev. E55, 4120 (1997)<br />

38. U.C. Täuber, J.E. Santos, and Z. Rácz: Eur. Phys. J. B7, 309 (1999)<br />

39. B. Schmittmann and R.K.P. Zia: Phys. Rev. Lett. 66, 357 (1991)<br />

40. B. Schmittmann: Europhys. Lett. 24, 109 (1993)<br />

41. K.E. Bassler and Z. Rácz: Phys. Rev. Lett. 73, 1320 (1994); Phys. Rev. E52,<br />

R9 (1995)<br />

42. B. Schmittmann and R.K.P. Zia: Statistical mechanics of driven diffusive systems.<br />

In: Phase Transitions and Critical Phenomena, vol. 17, ed by C. Domb<br />

and J.L. Lebowitz (Academic Press, London 1995)<br />

43. H.K. Janssen and B. Schmittmann: Z. Phys. B63, 517 (1986)<br />

44. K.-t. Leung and J.L. Cardy: J. Stat. Phys. 44, 567 (1986)<br />

45. D. Forster, D.R. Nelson, and M.J. Stephen: Phys. Rev. A16, 732 (1977)<br />

46. J.D. Murray: Mathematical Biology, vols. I/II (<strong>Spr<strong>in</strong>ger</strong>, New York, 3rd ed 2002)<br />

47. M. Mobilia, I.T. Georgiev, and U.C. Täuber: e-pr<strong>in</strong>t q-bio.PE/0508043 (2005)<br />

48. M. Doi: J. Phys. A: Math. Gen. 9, 1465 & 1479 (1976)<br />

49. P. Grassberger and M. Scheunert: Fortschr. Physik 28, 547 (1980)<br />

50. L. Peliti: J. Physique 46, 1469 (1985)<br />

51. F.C. Alcaraz, M. Droz, M. Henkel, and V. Rittenberg: Ann. of Phys. 230, 250<br />

(1994)<br />

52. M. Henkel, E. Orland<strong>in</strong>i, and J. Santos: Ann. of Phys. 259, 163 (1997)<br />

53. G.M. Schütz: In: Phase Transitions and Critical Phenomena, vol. 19, ed by<br />

C. Domb and J.L. Lebowitz (Academic Press, London 2001)<br />

54. R. St<strong>in</strong>chcombe: Adv. Phys. 50, 431 (2001)<br />

55. F. van Wijland: Phys. Rev. E63, 022101 (2001)<br />

56. J.W. Negele and H. Orland: Quantum many-particle systems (Addison-Wesley,<br />

Redwood City 1988)<br />

57. L. Peliti: J. Phys. A: Math. Gen. 19, L365 (1986)<br />

58. H.K. Janssen: J. Stat. Phys. 103, 801 (2001)<br />

59. B.P. Lee: J. Phys. A: Math. Gen. 27, 2633 (1994)<br />

60. B.P. Lee and J. Cardy: Phys. Rev. E50, R3287 (1994)


348 U. C. Täuber<br />

61. D. Toussa<strong>in</strong>t and F. Wilczek: J. Chem. Phys. 78, 2642 (1983)<br />

62. B.P. Lee and J. Cardy: J. Stat. Phys. 80, 971 (1995)<br />

63. O. Deloubrière, H.J. Hilhorst, and U.C. Täuber:Phys.Rev.Lett.89, 250601<br />

(2002); H.J. Hilhorst, O. Deloubrière, M.J. Washenberger, and U.C. Täuber: J.<br />

Phys. A: Math. Gen. 37, 7063 (2004)<br />

64. H.J. Hilhorst, M.J. Washenberger, and U.C. Täuber: J. Stat. Mech. P10002<br />

(2004)<br />

65. M. Moshe: Phys. Rep. 37C, 255 (1978)<br />

66. P. Grassberger and K. Sundermeyer: Phys. Lett. B77, 220 (1978)<br />

67. P. Grassberger and A. De La Torre: Ann. of Phys. 122, 373 (1979)<br />

68. H.K. Janssen and U.C. Täuber: Ann. of Phys. 315, 147 (2005)<br />

69. S.P. Obukhov: Physica A101, 145 (1980)<br />

70. J.L. Cardy and R.L. Sugar: J. Phys. A: Math. Gen. 13, L423 (1980)<br />

71. H.K. Janssen: Z. Phys. B42, 151 (1981)<br />

72. W. K<strong>in</strong>zel: In: Percolation structures and processes,edbyG.Deutsch,R.Zallen,<br />

and J. Adler (Hilger, Bristol 1983)<br />

73. P. Grassberger, Z. Phys. B47, 365 (1982)<br />

74. H. H<strong>in</strong>richsen: Adv. Phys. 49, 815 (2001)<br />

75. G. Ódor:Rev.Mod.Phys.76, 663 (2004)<br />

76. P. Grassberger: Math. Biosc. 63, 157 (1983)<br />

77. H.K. Janssen: Z. Phys. B58, 311 (1985)<br />

78. J.L. Cardy and P. Grassberger: J. Phys. A: Math. Gen. 18, L267 (1985)<br />

79. J. Benzoni and J.L. Cardy: J. Phys. A: Math. Gen. 17, 179 (1984)<br />

80. E. Frey, U.C. Täuber, and F. Schwabl: Europhys. Lett. 26, 413 (1994); Phys.<br />

Rev. E49, 5058 (1994)<br />

81. H.K. Janssen and O. Stenull: Phys. Rev. E62, 3173 (2000)<br />

82. U.C. Täuber, M.J. Howard, and H. H<strong>in</strong>richsen: Phys. Rev. Lett. 80, 2165<br />

(1998); Y.Y. Goldschmidt, H. H<strong>in</strong>richsen, M.J. Howard, and U.C. Täuber: Phys.<br />

Rev. E59, 6381 (1999)<br />

83. R. Kree, B. Schaub, and B. Schmittmann: Phys. Rev. A39, 2214 (1989)<br />

84. F. van Wijland, K. Oerd<strong>in</strong>g, and H. Hilhorst: Physica A251, 179 (1998);<br />

K. Oerd<strong>in</strong>g, F. van Wijland, J.P. Leroy, and H. Hilhorst: J. Stat. Phys. 99,<br />

1365 (2000)<br />

85. J. Cardy and U.C. Täuber: Phys. Rev. Lett. 77, 4780 (1996); J. Stat. Phys.<br />

90, 1 (1998)<br />

86. L. Canet, H. Chaté, and B. Delamotte: Phys. Rev. Lett. 92, 255703 (2004);<br />

L. Canet, H. Chaté, B. Delamotte, I. Dornic, and M.A. Muñoz: e-pr<strong>in</strong>t<br />

cond-mat/0505170 (2005)<br />

87. M. Henkel and H. H<strong>in</strong>richsen: J. Phys. A: Math. Gen. 37, R117 (2004)<br />

88. H.K. Janssen, F. van Wijland, O. Deloubrière, and U.C. Täuber: Phys. Rev.<br />

E70, 056114 (2004)<br />

89. H.K. Janssen: Phys. Rev. E55, 6253 (1997)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!